Register      Login
Journal of Southern Hemisphere Earth Systems Science Journal of Southern Hemisphere Earth Systems Science SocietyJournal of Southern Hemisphere Earth Systems Science Society
A journal for meteorology, climate, oceanography, hydrology and space weather focused on the southern hemisphere
RESEARCH ARTICLE (Open Access)

ACCESS datasets for CMIP6: methodology and idealised experiments

C. Mackallah https://orcid.org/0000-0003-4989-5530 A * , M. A. Chamberlain https://orcid.org/0000-0002-3287-3282 B , R. M. Law https://orcid.org/0000-0002-7346-0927 A , M. Dix https://orcid.org/0000-0002-7534-0654 A , T. Ziehn https://orcid.org/0000-0001-9873-9775 A , D. Bi A , R. Bodman https://orcid.org/0000-0002-8349-3001 A C , J. R. Brown https://orcid.org/0000-0002-1100-7457 C D , P. Dobrohotoff https://orcid.org/0000-0001-7315-042X A , K. Druken E , B. Evans E , I. N. Harman https://orcid.org/0000-0002-5690-0484 F , H. Hayashida https://orcid.org/0000-0002-6349-4947 G D , R. Holmes https://orcid.org/0000-0002-6799-9109 H , A. E. Kiss https://orcid.org/0000-0001-8960-9557 I D , A. Lenton https://orcid.org/0000-0001-9437-8896 B , Y. Liu E , S. Marsland https://orcid.org/0000-0002-5664-5276 A , K. Meissner https://orcid.org/0000-0002-0716-7415 J D , L. Menviel https://orcid.org/0000-0002-5068-1591 J K , S. O’Farrell https://orcid.org/0000-0002-9019-6136 A , H. A. Rashid https://orcid.org/0000-0003-1896-2446 A , S. Ridzwan E , A. Savita https://orcid.org/0000-0003-2800-3636 A D G , J. Srbinovsky A , A. Sullivan https://orcid.org/0000-0002-5712-6195 A , C. Trenham https://orcid.org/0000-0003-4258-9936 F , P. F. Vohralik L , Y.-P. Wang https://orcid.org/0000-0002-4614-6203 A , G. Williams https://orcid.org/0000-0002-2805-7426 M , M. T. Woodhouse https://orcid.org/0000-0002-9892-4492 A and N. Yeung https://orcid.org/0000-0002-6560-6658 D J
+ Author Affiliations
- Author Affiliations

A CSIRO Oceans and Atmosphere, Aspendale, Vic. 3195, Australia.

B CSIRO Oceans and Atmosphere, Battery Point, Tas. 7004, Australia.

C School of Geography, Earth and Atmospheric Sciences, The University of Melbourne, Parkville, Vic. 3010, Australia.

D Australian Research Council Centre of Excellence for Climate Extremes (CLEX), Sydney, NSW 2052, Australia.

E National Computational Infrastructure, Acton, ACT 2601, Australia.

F CSIRO Oceans and Atmosphere, Black Mountain, ACT 2601, Australia.

G Institute for Marine and Antarctic Studies, University of Tasmania, Hobart, Tas. 7005, Australia.

H School of Geosciences, University of Sydney, Camperdown, NSW 2006, Australia.

I Research School of Earth Sciences, Australian National University, Acton, ACT 2601, Australia.

J Climate Change Research Centre, University of New South Wales, Sydney, NSW 2052, Australia.

K The Australian Centre for Excellence in Antarctic Science, University of Tasmania, Hobart, Tas. 7001, Australia.

L CSIRO Oceans and Atmosphere, Lindfield, NSW 2070, Australia.

M CSIRO Information Management & Technology, Black Mountain, ACT 2601, Australia.

* Correspondence to: chloe.mackallah@csiro.au

Journal of Southern Hemisphere Earth Systems Science 72(2) 93-116 https://doi.org/10.1071/ES21031
Submitted: 16 December 2021  Accepted: 26 May 2022   Published: 14 July 2022

© 2022 The Author(s) (or their employer(s)). Published by CSIRO Publishing on behalf of BoM. This is an open access article distributed under the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License (CC BY-NC-ND)

Abstract

The Australian Community Climate and Earth System Simulator (ACCESS) has contributed to the World Climate Research Programme’s Coupled Model Intercomparison Project Phase 6 (CMIP6) using two fully coupled model versions (ACCESS-CM2 and ACCESS-ESM1.5) and two ocean–sea-ice model versions (1° and 0.25° resolution versions of ACCESS-OM2). The fully coupled models differ primarily in the configuration and version of their atmosphere components (including the aerosol scheme), with smaller differences in their sea-ice and land model versions. Additionally, ACCESS-ESM1.5 includes biogeochemistry in the land and ocean components and can be run with an interactive carbon cycle. CMIP6 comprises core experiments and associated thematic Model Intercomparison Projects (MIPs). This paper provides an overview of the CMIP6 submission, including the methods used for the preparation of input forcing datasets and the post-processing of model output, along with a comprehensive list of experiments performed, detailing their initialisation, duration, ensemble number and computational cost. A small selection of model output is presented, focusing on idealised experiments and their variants at global scale. Differences in the climate simulation of the two coupled models are highlighted. ACCESS-CM2 produces a larger equilibrium climate sensitivity (4.7°C) than ACCESS-ESM1.5 (3.9°C), likely a result of updated atmospheric parameterisation in recent versions of the atmospheric component of ACCESS-CM2. The idealised experiments run with ACCESS-ESM1.5 show that land and ocean carbon fluxes respond to both changing atmospheric CO2 and to changing temperature. ACCESS data submitted to CMIP6 are available from the Earth System Grid Federation (https://doi.org/10.22033/ESGF/CMIP6.2281 and https://doi.org/10.22033/ESGF/CMIP6.2288). The information provided in this paper should facilitate easier use of these significant datasets by the broader climate community.

Keywords: ACCESS, climate change, climate data, climate model evaluation, climate simulation, CMIP6, coupled climate model, Earth System Model.

1. Introduction

The Coupled Model Intercomparison Project (CMIP), overseen by the World Climate Research Programme (WCRP), endeavours to collate the results of a standardised set of experiments from global climate models developed at climate research organisations across the globe. The coordinated experiment design and implementation, along with the consolidated results of many different global climate models, allows for robust analysis and a deep investigation of the physical processes of the climate system and the biases inherent in numerical climate models. The current phase, CMIP6, consists of a base set of idealised experiments (known as the DECK; Eyring et al. 2016) and historical simulations that modelling centres are required to perform in order to participate in CMIP6. In addition, ~322 different experiments have been designed across the 23 Model Intercomparison Projects (MIPs) endorsed by the World Climate Research Programme1 (WCRP). These MIPs address a range of investigatory avenues – for example, exploring future climate projections under a range of socioeconomic scenarios (ScenarioMIP; O’Neill et al. 2016), and investigating the effects of removing carbon dioxide (CO2) from the atmosphere (CDRMIP; Keller et al. 2018).

Since the conclusion of CMIP5, two new iterations of the Australian Community Climate and Earth System Simulator (ACCESS), focusing on global climate, have been developed for the submission of simulation data to CMIP6. ACCESS-CM2 (Bi et al. 2020) simulates the physical climate with an updated atmospheric component, whereas ACCESS-ESM1.5 (Ziehn et al. 2020a) has been designed to simulate a fully-interactive carbon cycle; both models also have updated ocean and sea-ice components compared to the CMIP5-era versions of ACCESS. Although the two ACCESS models use broadly similar modelling components, ACCESS-CM2 uses newer versions of the atmospheric, land and sea-ice components, whereas ACCESS-ESM1.5 contains additional land and ocean biogeochemical components to facilitate the carbon cycle. Additionally, an ocean and sea-ice-only version of ACCESS-CM2 has been developed in partnership with the Consortium for Ocean–Sea-Ice Modelling in Australia (COSIMA) at three resolutions – ACCESS-OM2 (1° ocean), ACCESS-OM2-025 (0.25° ocean) and ACCESS-OM2-01 (0.1° ocean, which was not submitted to CMIP6), all referred to collectively as ‘ACCESS-OM2’ (Kiss et al. 2020). See Table 1 for details on the modelling components of each version of ACCESS. ACCESS has participated in the core CMIP/DECK and nine other MIPs, according to the needs of the Australian modelling community and the strengths of each model version.


Table 1.  Summary of modelling components for the four ACCESS global climate models submitted to CMIP6.
Click to zoom

One purpose of this paper is to document key details of the ACCESS submissions to CMIP6. Although there are well-defined protocols for each CMIP6 experiment, the differing complexities of participating models and resources available to each modelling group mean that different choices (and sometimes compromises) are made as to how to meet the CMIP6 experimental specifications. Hence, one component of this paper (Section 1) documents these choices for the ACCESS models, including the spin-up that was performed before experiments were commenced and how the various forcing datasets were applied to each model configuration. A summary of the differences in model versions is provided, along with details of relevant model description and evaluation papers. Section 1 also provides a summary of all the experiments performed for CMIP6 with each ACCESS model, and the ACCESS datasets that are currently available on the Earth System Grid Federation (ESGF, which hosts the CMIP6 submissions across a number of data centres; Balaji et al. 2018). A description of the post-processing pipeline used to conform the model output and metadata to CMIP standards is provided, along with a brief analysis of the computational costs associated with the CMIP6 effort of our modelling group using prescribed metrics. The aim of this section is to provide sufficient information to support the reproducibility of experiments and further analysis of the existing ACCESS datasets.

A second aim of this paper (Section 2) is to provide a brief overview of some of the output from the DECK idealised experiments and related experiments from C4MIP (Jones et al. 2016) and CDRMIP (Keller et al. 2018). These experiments are used to characterise key features of a model’s behaviour such as climate sensitivity, climate drift and overall performance against expected climate responses. The analysis focuses on global mean responses and notes how the simulated climate varies across the two ACCESS coupled models. Global carbon fluxes and related biogeochemical variables are presented for ACCESS-ESM1.5 simulations, with a focus on how changing atmospheric CO2 and changing temperature affect the behaviour of the carbon cycle. This overview is intended as a springboard and to provide context for further studies, and to compare ACCESS simulations with other CMIP6 models in cases where multimodel studies have already been published.

The paper concludes in Section 3 with some comments on the impact of the ACCESS submissions to CMIP6, whereas Section 3 directs the reader to repositories where the model codes and post-processed data are available.


2. Implementation of the CMIP6 protocol for ACCESS experiments

This section documents the steps required for the ACCESS models to participate in CMIP6, including a summary of the model configurations used, the experiments performed, the application of the forcing datasets for use with the ACCESS model and the post-processing required to meet CMIP6 standards.

2.1. Models

For its atmospheric component, ACCESS-CM2 uses the UK Met Office (UKMO) Unified Model (ver. 10.6, UM) GA7.1 configuration (Walters et al. 2019, MetUM-HadGEM3-GA7.1) with a vertical resolution of 85 levels and the GLOMAP-mode aerosol scheme (Mann et al. 2010). This is the same atmospheric configuration and resolution as the HadGEM3-GC31-LL and KACE-1-0-G CMIP6 submissions. ACCESS-ESM1.5 is based on an updated version of ACCESS1.3 (which was submitted to CMIP5; see Bi et al. 2013), and uses a UM (ver. 7.3) configuration that is close to GA1 (Bellouin et al. 2011a, HadGAM2) with 38 levels and the CLASSIC aerosol scheme (Bellouin et al. 2011b). Both are run at ‘N96’ horizontal resolution (1.875 longitude by 1.25 latitude) and make use of the Australian-developed land surface scheme CABLE (Community Atmosphere Biosphere Land Exchange; Kowalczyk et al. 2006, 2013), which is integrated into the UM atmosphere. The terrestrial carbon cycle is implemented in ACCESS-ESM1.5 using CASA-CNP (Wang et al. 2010), within CABLE, including accounting for the impact of nitrogen and phosphorus on the carbon cycle.

Additional components of ACCESS include the LANL CICE model for the sea-ice component (Hunke and Lipscomb 2010), and the NOAA/GFDL Modular Ocean Model (ver. 5, MOM5) for the ocean component at nominal 1° resolution (Griffies 2012). CICE and MOM5 are also the primary components of ACCESS-OM2, with 0.25° version of CICE and MOM5 used in ACCESS-OM2-025. ACCESS-ESM1.5 and ACCESS-OM2 (1° version only) also include the ocean biogeochemical modelling component WOMBAT (Whole Ocean Model of Biogeochemistry And Trophic-dynamics, Oke et al. 2013; Law et al. 2017). A tabulated breakdown of the differences between the ACCESS models used for CMIP6 can be found in Table 1.

More extensive documentation is available elsewhere. Bi et al. (2020) describes the ACCESS-CM2 configuration and the spin-up of the model. Climate drift is assessed and the model performance is evaluated including the mean state of the ocean and aspects of climate variability. The simulated climate is compared with that of ACCESS1.3, as used in CMIP5. The climate of this model version has also been evaluated for an atmosphere-only (amip) simulation by Bodman et al. (2020), looking at both global and Australian metrics. Ziehn et al. (2020a) describes the ACCESS-ESM1.5 configuration including its performance compared with ACCESS-ESM1 (Law et al. 2017). The model is evaluated for climate and carbon cycle stability in the pre-industrial control simulation, whereas the later part of the historical simulation is used to evaluate against observations. Kiss et al. (2020) describes the ACCESS-OM2 configuration and provides extensive evaluation of simulations at three different resolutions. The ACCESS-OM2 configurations used for CMIP6 were updated relative to those described in Kiss et al. (2020) including, most notably, improved topography and updated forcing from JRA55-do ver. 1.3 to ver. 1.4. When WOMBAT was active in ACCESS-OM2, the biogeochemical parameters were identical to those described by Ziehn et al. (2020a) and used in ACCESS-ESM1.5.

2.2. Experiments and ensemble methodology

2.2.1. Spin-up, DECK and historical experiments

Prior to beginning an official CMIP6 pre-industrial control experiment (piControl), a model must be initialised and spun-up under CMIP6 pre-industrial forcing conditions, in order to bring the model climate as closely into balance with the forcing as possible. The spin-up period of ACCESS-CM2 was 950 simulation years, during which several changes were made to the model, including the implementation of CABLE over the original land scheme JULES (Joint UK Land Environment Simulator; Best et al. 2011; Clark et al. 2011) and several bug fixes and tuning changes. The last 100 years of the ACCESS-CM2 spin-up were simulated using the final configuration of the model. The physical climate of ACCESS-ESM1.5 was spun-up over a period of 3000 simulation years with only minor changes and bug fixes applied during this time (with only very minor impacts on the climate trajectory), whereas the biogeochemical processes were integrated over the latter 1000 years of this spin-up. The last 600 years of spin-up represent the final ACCESS-ESM1.5 model configuration. OMIP-2 protocols do not require a spin-up period for the standard ocean-only experiments. Detailed descriptions of code changes and parameter tunings applied to the ACCESS models during their relative spin-up periods can be found in their respective model description papers: Bi et al. (2020) (ACCESS-CM2), Ziehn et al. (2020a) (ACCESS-ESM1.5) and Kiss et al. (2020) (ACCESS-OM2 and ACCESS-OM2-025).

The piControl experiment continues directly from the spin-up, with many experiments then branching from the first year of the piControl – designated year 0950 in ACCESS-CM2, and year 0101 in ACCESS-ESM1.5. The piControl underpins most other experiments by providing a baseline with which to account for residual drift in the climate system that is not part of its response to forcings. The piControl experiment is part of the DECK, which also includes an idealised 1% per year CO2 increase simulation (1pctCO2), an idealised simulation with an abrupt quadrupling of CO2 concentration (abrupt-4xCO2), an Atmospheric Model Intercomparison Project simulation with prescribed sea surface temperatures (amip) and a historical simulation over the period 1850–2014. Furthermore, for Earth System Models2 there is an additional piControl experiment and historical simulation required in which CO2 emissions are prescribed, and a fully interactive carbon cycle simulated. In these experiments, esm-piControl and esm-hist, atmospheric CO2 concentrations are calculated according to CO2 fluxes between the atmosphere, ocean and land; as opposed to prescribing atmospheric CO2 concentrations as in the piControl and historical.

2.2.2. Experiment characteristics and ensemble methodology

In addition to the DECK experiments, each MIP consists of Tier 1 (priority) and Tier 2 experiments. Each modelling group has chosen which experiments to complete with Tables 24 detailing the list of experiments performed for CMIP6 (as of April 2022) using ACCESS-CM2, ACCESS-ESM1.5 and ACCESS-OM2 respectively. The experiments are presented by MIP, with the start year referring to the first internal model year of the simulation, which is set to real dates where appropriate (e.g. in the historical simulation). For branched experiments that comprise only a single ensemble member, the branch year is usually the same as its start year (e.g. the ACCESS-ESM1.5 experiment 1pctCO2-cdr begins at year 0241, having branched from 1pctCO2 at the end of year 0240), except where the internal year is set to a real date after branching (most commonly in the case of historical-based simulations; e.g. the ACCESS-ESM1.5 experiment hist-bgc branches from the piControl at year 0161, but the internal start date is set to 1850).


Table 2.  ACCESS-CM2 (CSIRO-ARCCSS) experiments.
Click to zoom


Table 3.  ACCESS-ESM1.5 (CSIRO) experiments.
Click to zoom


Table 4.  ACCESS-OM2 (CSIRO-COSIMA) experiments.
Click to zoom

For experiments with multiple ensemble members, there are two different avenues of ensemble methodology: (i) new ensembles branching from a piControl or esm-piControl, and (ii) child ensembles branching from an existing parent ensemble. For case (i) – new ensembles that branch from a piControl (typically historical simulations) – the following methodology applies: each ensemble member of a child experiment is branched from a progressively advanced year of the piControl (or esm-piControl), and the ‘variant label’ of the child member is advanced by one ‘realisation’ (e.g. r1i1p1f1 and r2i1p1f1 are realisations one and two respectively; see the CMIP6 Controlled Vocabulary3 for details). The branch years of new ensembles are specified, where appropriate, in the right-hand column of Tables 2, 3; branch years are 50 years apart for ACCESS-CM2 and 20 years apart for ACCESS-ESM1.5. In cases where avenue (ii) applies, ensemble members of the child experiment will match those of its parent – for example, ensemble member three (r3i1p1f1) of ssp126 was branched from ensemble member three (r3i1p1f1) of the historical. In these cases – primarily future scenario projections that branch from historical simulations – the ‘variant label’ of the child remains the same as the parent from which it was branched.

A second ACCESS-ESM1.5 abrupt-4xCO2 simulation (r2i1p1f1) was run for 1000 years, for submission to the ongoing LongRunMIP4 (Rugenstein et al. 2019), which is not an official CMIP activity. Furthermore, a large ensemble (defined as >10 realisations by Deser et al. (2020)) consisting of 40 members has been simulated with ACCESS-ESM1.5 for the historical and tier 1 ScenarioMIP experiments (ssp126, ssp245, ssp370 and ssp585), along with a 30-member ensemble for the four experiments that were simulated from the CovidMIP activity (which sits within DAMIP and comprises a total of six experiments; Gillett et al. 2016; Jones et al. 2021). These large ensembles are available for use by the SMILE5 (Single Model Initial-condition Large Ensemble) community, and will be investigated in more detail in papers currently in preparation. Work is also ongoing to expand the ACCESS-CM2 historical and ScenarioMIP ensembles, with the intention to produce a 10-member ensemble.

Three ACCESS-OM2 experiments (see Table 4) were contributed to OMIP through the OMIP-2 protocol (Griffies et al. 2016; Orr et al. 2017; Tsujino et al. 2020) forced by repeated 61-year (1958–2018) cycles of the Japanese atmospheric reanalysis (JRA55-do, ver. 1.4; Tsujino et al. 2018). The 1° ACCESS-OM2, including the ocean biogeochemistry component WOMBAT, ran both OMIP-2 experiments, omip2 (six cycles) and omip2-spunup (six cycles after conducting 33 cycles as spin-up). The 0.25° ACCESS-OM2-025, without the ocean biogeochemistry component, was also used to run the omip2 experiment (six cycles).

2.3. Forcings

Depending on the experiment, ACCESS requires time-varying forcings, including prescribed monthly solar forcing, greenhouse gases (GHGs), volcanic aerosol optical depth, aerosol chemistry emissions and ozone (Eyring et al. 2016). Land-use and land-cover forcing variations have also been implemented for ACCESS-ESM1.5. Most of the input forcing data were supplied either directly from CMIP6 by input4MIPs6 (Meinshausen and Vogel 2016; Durack et al. 2018) or prepared by the UKMO for use in their HadGEM3-GC31-LL CMIP6 submission. Other forcings were determined as specified below.

Idealised experiments in CMIP6, such as 1pctCO2 and abrupt-4xCO2, largely follow the piControl, with historical data used in the historical and related experiments (such as DAMIP; Gillett et al. 2016) and the amip simulation. Other MIPs provide or specify additional forcing data (e.g. ScenarioMIP; O’Neill et al. 2016) or protocols (e.g. RFMIP; Pincus et al. 2016).

In general, we adopted the same forcing files for ACCESS-CM2 as were used for HadGEM3-GC3.1 (Sellar et al. 2020), whereas ACCESS-ESM1.5 required further alterations to input data, including interpolation to the HadGAM2 grid, or the use of separately prepared ancillary data as with land-use change forcings. The forcings used for the PMIP (Kageyama et al. 2018) experiments performed with the ACCESS-ESM1.5 (lig127k and midHolocene) follow the PMIP4 protocols (Otto-Bliesner et al. 2017).

2.3.1. Solar variability

For ACCESS-CM2, piControl (and associated experiments) solar forcing is based on a mean total solar insolation (TSI) value of 1361.0 W m−2, combined with a monthly time-varying spectral file that provides details for the solar cycle, based on the time-averaged mean conditions of the two solar cycles covering the years 1850–1873, and is derived from the recommended solar datasets for CMIP6 (Matthes et al. 2017).

For ACCESS-ESM1.5, the setup was slightly different. The model spin up was initiated prior to release of CMIP6 protocols, therefore both the spin-up and all idealised experiments, including the piControl were performed using the CMIP5 solar constant of 1365.65 W m−2, as it was in the CMIP5 versions of ACCESS. ACCESS-ESM1.5 does not utilise full spectral variations.

Historical and future scenario solar variability was based on the historical reconstructions of the 1850–2014 period, and projections of the 2015–2299 period, by Matthes et al. (2017). For ACCESS-ESM1.5, the CMIP6 TSI data were adjusted by a constant 4.9 W m−2 offset in the historical and future simulations to match the CMIP5-style piControl setup (Ziehn et al. 2020a).

All experiments except PMIP use orbital parameters corresponding to the year 2000. For PMIP experiments, Earth’s orbital parameters (eccentricity, longitude of perihelion and obliquity) were adjusted to the estimated values of the Last Interglacial and mid-Holocene periods (Berger and Loutre 1991; Otto-Bliesner et al. 2017). These orbital parameters modulate the magnitude, and seasonal and latitudinal distribution of the incoming solar radiation.

2.3.2. Stratospheric volcanic aerosol

Stratospheric volcanic aerosol optical depths (AODs), originating from explosive volcanic events, are determined from the zonal mean data of Arfeuille et al. (2014) and Thomason et al. (2018), with aerosol optical properties in the solar (shortwave) and terrestrial (longwave) spectrum. As per the CMIP6 protocols, a climatology of the 1850–2014 period was used for all idealised experiments and future scenarios past 2025, with a smooth transition occurring in 2015–2024 between the 2014 values and the climatology.

ACCESS-ESM1.5 and ACCESS-CM2 use different approaches for specifying stratospheric aerosols, with ACCESS-CM2 utilising AODs as a function of latitude, height and wavelength, as processed by the Easy Volcanic Aerosol module (Toohey et al. 2016). For ACCESS-ESM1.5, we derived AODs over four equal latitude bands from the simple 550 nm AOD data provided for use in CMIP6 (Ziehn et al. 2020a). A small offset was applied to account for the difference in the average AOD from CMIP6 and the CMIP5 data used for the ACCESS-ESM1-5 spin-up and piControl.

2.3.3. Well-mixed GHGs

The GHG concentrations are included in ACCESS models as globally uniform annual mass mixing ratios. ACCESS-CM2 uses CO2, methane (CH4), nitrous oxide (N2O), CFC12-eq and HFC134a-eq, with the latter two representing groups of chlorofluorocarbons (CFCs) and hydrofluorocarbons (HFCs) (Eyring et al. 2016). ACCESS-ESM1.5 uses CO2, CH4, N2O and separate values for CFC11, CFC12, CFC113, HCFC22, HFC125 and HFC134a, which together account for 98% of the change in the historical radiative forcing (Meinshausen et al. 2017). In the piControl, these values follow: CO2 = 284.317 ppm, CH4 = 808.25 ppb and N2O = 273.02 ppb. The GHG concentrations from Meinshausen et al. (2017) were used for the historical and related experiments (such as amip), whereas for all future scenarios the values from Meinshausen et al. (2020) were used. For PMIP experiments, the concentrations of CO2, CH4 and N2O were adjusted to the estimated concentrations of the mid-Holocene and Last Interglacial (lig127k: CO2 = 275 ppm, CH4 = 685 ppb and N2O = 255 ppb; midHolocene: CO2 = 264.4 ppm, CH4 = 597 ppb and N2O = 262 ppb; Otto-Bliesner et al. 2017).

ACCESS-ESM1.5 was also used for CO2 emissions-driven simulations (such as esm-historical) in interactive mode, with emissions prescribed by Hoesly et al. (2018). Emissions sources were combined and released from the surface, following interpolation to the ACCESS-ESM1.5 grid and scaling to ensure that total annual emissions agree with the CEDS project.7 All other GHG concentrations are prescribed for ACCESS-ESM1.5 emissions-driven simulations.

2.3.4. Aerosol emissions (anthropogenic and natural)

Both ACCESS-CM2 and ACCESS-ESM1.5 use emissions of black carbon, organic carbon, sulfur dioxide, dimethyl sulfide (DMS) and sea salt for determining optical depths from tropospheric aerosol. Sources of carbon emissions include anthropogenic fossil-fuel and biofuel burning (Hoesly et al. 2018), as well as the forest biomass burning emissions of van Marle et al. (2017), as per CMIP6 protocols. Sulfur dioxide emissions originate from both prescribed anthropogenic (Hoesly et al. 2018) and natural (volcanic degassing, calculated internally by climatology) sources which act as a precursor to secondary sulfate aerosol, along with DMS emissions, which are simulated interactively using prescribed seawater DMS. Sea salt emissions are also simulated interactively in both models. Biogenic and natural emissions are not supplied by CMIP6, and are implemented as in the setup of the UKMO CMIP6 models (Sellar et al. 2020), including a fixed monthly climatology of terpene emissions. Mineral dust is not included in the aerosol schemes of either model, and is simulated separately using the six-bin mass-based scheme of Woodward (2011).

In ACCESS-CM2, aerosol emissions are processed, and their evolution and interactions simulated, by the UKCA GLOMAP-mode module (Mann et al. 2010); whereas ACCESS-ESM1.5 uses the CLASSIC scheme of Bellouin et al. (2011b). Emission files were adopted from the setup of the UKMO GC3.1 model simulations, and were used directly in ACCESS-CM2, but ACCESS-ESM1.5 required regridding prior to use.

2.3.5. Ozone

Although some CMIP6 models simulate ozone, many prescribe ozone distributions. Both ACCESS configurations apply prescribed monthly ozone concentrations with ACCESS-CM2 using 3-D fields and ACCESS-ESM1.5 using zonal means. This input data originated from the ozone data sets of Morgenstern et al. (2017) for stratospheric ozone, and Checa-Garcia et al. (2018) for tropospheric ozone, which were interpolated from pressure levels to hybrid height coordinates by the UKMO (Sellar et al. 2020). For ACCESS-ESM1.5, further interpolation from 85 to 38 model levels was required.

In ACCESS-CM2 simulations (except for piControl), an additional calculation was applied at each simulation year in order to ensure that the prescribed ozone and model tropopause remain consistent as the tropopause height changes with temperature. The ozone redistribution scheme of Hardiman et al. (2019) redistributes excess tropospheric ozone concentrations into the stratosphere, while conserving global ozone mass. ACCESS-ESM1.5 has much lower vertical resolution about the tropopause and this adjustment would have much less effect and is not included.

2.3.6. Land use change and nitrogen deposition

ACCESS-ESM1.5 uses a simple land-use scheme which allows for annual changes in tile fractions, and reallocates carbon, nitrogen and phosphorus pools accordingly (Zhang et al. 2013; Ziehn et al. 2020a). Forcing data for changes in vegetation fractions were derived from the Land-Use Harmonisation 2 (LUH2) dataset (Hurtt et al. 2017), which were mapped onto the CABLE plant functional types used in the model.

Nitrogen and phosphorus deposition forcings are used in all ACCESS-ESM1.5 simulations. Nitrogen deposition data are provided by CMIP6 (Jones et al. 2016), whereas phosphorus deposition data (not provided in CMIP6) were adopted from simulations of the predecessor of ACCESS-ESM1.5, ACCESS-ESM1, based on datasets used in Wang et al. 2010).

2.3.7. Atmosphere-only sea surface temperature and sea-ice cover forcings

Sea surface temperatures (SSTs) and sea-ice concentrations are prescribed in the atmosphere-only amip simulations of the DECK, the data for which were derived from the input4MIPs datasets of Durack and Taylor (2017), and interpolated to the UM model grid and N96 land-sea fractional mask by the UKMO.

These data were used in all ACCESS-ESM1.5 amip simulations, and in the first three ACCESS-CM2 amip ensemble members (r[1,2,3]i1p1f1), after which time it was discovered that a mismatch between the ACCESS-CM2 land-sea mask and that used to create the ancillary input files resulted in undefined SSTs at some coastal points, for which the model used a default SST of 273.1 K.8 A corrected set of SST and sea-ice ancillary files were used for an additional set of three amip simulations with ACCESS-CM2, which were assigned the variant strings r[1,2,3]i1p1f2. We note here that the amip evaluation by Bodman et al. (2020) included only the initial set of ACCESS-CM2 amip ensemble members (r[1,2,3]i1p1f1). ACCESS-ESM1.5 simulations were not affected.

The SST and sea-ice concentrations were also prescribed in all RFMIP experiments. Climatologies were calculated from the first 50 years of the piControl experiment of each model. These were then interpolated to daily values, and adjusted following the AMIP II boundary condition calculation of Taylor et al. (2000) to ensure that the interpolated monthly means match the original climatology. No issues arose from these ancillary files, because they were internally consistent with each of the models and their associated land-sea masks.

2.4. Post-processing

The creation of datasets suitable for publication to the ESGF requires significant resources. An outline of the process for ACCESS datasets is provided here. The raw model output from the UM atmospheric component (along with the CABLE land component output) of ACCESS exists in a binary file format (known as ‘Fieldsfiles’ or the ‘PP-format’9); and is converted to netCDF4 using the Python package Iris (Met Office 2020) as a first step in the post-processing pipeline (noting that the first several CMIP6 production simulations used cdms210 instead). All other components of ACCESS natively produce model output in netCDF format.

A Python-based software package, referred to as the ‘ACCESS Post-Processor ver. 4.0’ (APP4; Mackallah et al. 2022), has been developed to enable the generation of CMIP6 data using ACCESS model output, and is based on an earlier version of the APP that was used for the ACCESS submission to CMIP5 (Collier and Uhe 2012; Uhe et al. 2012). APP4 uses a python2.7 environment and cdms2 to extract and prepare the relevant fields from the model output, and relies on PCMDI’s ‘Climate Model Output Rewriter’ (CMOR ver. 3.411; Nadeau et al. 2016) to create final data products, ensuring that they adhere to CMIP6 quality standards and follow the CMIP6 data reference syntax12 (DRS) and CF metadata conventions.13 Note that the model_id for ACCESS-ESM1.5 is ACCESS-ESM1-5, this is reflected in all directory and file names. Primary inputs include an experiment-specific data request (created using the dreqPy tool ver. 01.00.31, Juckes et al. 2020), the CMIP6 CMOR Tables,14 which define the controlled vocabulary and conventions, and an internally created mapping file that specifies how each CMIP6 variable is generated from the existing fields in the ACCESS model output. Reproducibility and code availability are detailed in Section 3.

Key variables that require significant treatment beyond simple arithmetical calculations include overturning circulations and streamfunctions (e.g. msftyz), ocean and sea-ice transports (e.g. mfo), and sea-ice extents, volumes and areas (e.g. siextentn). These are performed by APP4 internally with Python functions, prior to the use of CMOR. Furthermore, many variables require minor manipulations such as averaging over spatial, temporal or tiled dimensions (e.g. most annual variables are based on monthly data), unit conversions (performed by CMOR where possible), masking or the filling of missing values, and the calculation of climatologies. Basic quality checking of metadata is performed using CMOR’s PrePARE tool, along with automatically generated plots to identify spurious and missing data.

As of 11 April 2022, 37 028 ACCESS-CM2 datasets, 265 166 ACCESS-ESM1.515 datasets and 822 ACCESS-OM2 datasets have been submitted to the ESGF,16 where each dataset is a single version of an individual variable at a given time frequency for each ensemble member. Version strings (e.g. ver. 20191225) contain the date on which the dataset was generated, with new versions being generated in the case of retraction and republication (see the CMIP Errata service for details on all ACCESS dataset retractions17). There have been DOIs minted for datasets grouped by MIP and the relevant citations are noted in Tables 24.

2.5. Computational costs and CPMIP metrics

All simulations with ACCESS models for CMIP6 were performed on the High Performance Computing (HPC) systems of the National Computational Infrastructure (NCI) at the Australian National University. The supercomputer Raijin was used for the first several ACCESS simulations performed, including the piControl-spinup and most of the DECK experiments; and for the historical and ScenarioMIP experiments, the first realisation of ACCESS-CM2 and first three realisations of ACCESS-ESM1.5. All later simulations were run on its replacement system Gadi. Tests with ensembles of fifty 1-year simulations showed that the climates simulated on the two machines were indistinguishable. Although not bit-wise reproducible, the results were consistent across all variables to the same level of internal climate variability seen among ensemble members, indicating that the model simulations were consistent across HPC systems.

ACCESS-CM2 simulations were performed in suites developed using the Rose engine18 and scheduled using the Cylc workflow engine.19 Most ACCESS-ESM1.5 simulations were configured and scheduled with scripts held at NCI, with the exception of PMIP experiments which were run using the payu framework developed by the ARC Centre of Excellence for Climate Extremes (CLEX). The payu framework was also used for the ACCESS-OM2 simulations. Code availability is detailed in Section 5.

Following an assessment of the computational performance of CMIP6 Earth System Models by Balaji et al. (2017), we present the performance of the ACCESS models against the metrics described by the Computational Performance MIP (CPMIP) project in Table 5. Values are taken from simulations performed on NCI’s Gadi HPC system (Cascade Lake nodes). The CPMIP resolution is the total number of 3-D grid points, and in the coupled model configurations is dominated by the ocean, so ACCESS-CM2 and ACCESS-ESM1.5 have quite similar values. However, the extra complexity of the atmospheric physics means that the atmosphere dominates the computational cost and limits the scaling in ACCESS-CM2. Therefore there is no close relation between resolution and cost in these metrics. Complexity (number of prognostic variables) is larger for ACCESS-CM2 than ACCESS-ESM1.5 because of the more sophisticated atmospheric chemistry and aerosol scheme. It is larger for ACCESS-OM2 compared to ACCESS-OM2-025 because only the former includes the ocean biogeochemical component WOMBAT. For ACCESS-CM2 and ACCESS-ESM1.5, data intensity (measured in gigabytes per compute hour) and data output cost depend on whether sub-daily data are saved from the atmospheric component; values presented are from a simulation in which sub-daily data were not saved.


Table 5.  Computational performance of ACCESS models following CPMIP (Balaji et al. 2017) metrics.
Click to zoom


3. Results from idealised experiments

In this section we present a range of metrics from different components of the climate system, both physical and biogeochemical, from the two fully coupled models ACCESS-CM2 and ACCESS-ESM1.5. The ocean-only ACCESS-OM2 results will be presented in a separate paper, in preparation. These metrics show the stability of the models in control experiments and also the responses to perturbations over a span of timescales in idealised climate warming scenarios. The figures that follow demonstrate that the impacts of these perturbations vary widely across the climate system; some components can be sensitive to a particular perturbation while insensitive to others. Some impacts are persistent but others are relatively transient. There is insufficient space here for a detailed study of the simulated fields or their responses to climate change. Rather, the overview here provides context for and motivates further analysis. The mean climate states and drifts in the models are already reported in their respective model description papers (Bi et al. 2020; Ziehn et al. 2020a), so the focus here is on comparing their behaviour. Climate responses from historical and future projections from ACCESS models are not covered here but where there is relevant evaluation as part of a multimodel study, this is cited in the discussions below.

This overview uses only a small selection of the variables submitted to CMIP6 (e.g. Dix et al. 2019a; Ziehn et al. 2019c) and their descriptions are listed in Table 6. The global mean physical climate variables used for ACCESS-CM2 and ACCESS-ESM1.5 are near-surface air temperature (tas), top of atmosphere (TOA) net energy balance down (rsdt − rsut − rlut), sea-ice area (siarean + siareas) and ocean temperature (thetaoga). For ACCESS-ESM1.5, the global biogeochemical (BGC) metrics used are net land carbon flux (nbp, positive into land), land net primary productivity (npp), net ocean carbon flux (fgco2, positive into ocean), ocean productivity (intpp), ocean O2 flux (fgo2, positive into ocean), ocean surface acidity (derived from dissic, talk and equilibrium constants) and global atmospheric CO2 mixing ratio (co23D).


Table 6.  List of CMIP6 variables used in the present analysis.
Click to zoom

We first discuss the control simulations and then the climate response for the experiments where CO2 is quadrupled (abrupt-4xCO2) or increases by 1% per year (1pctCO2). Effective radiative forcings calculated using RFMIP experiments (Pincus et al. 2016) are also presented. This is followed by C4MIP (Jones et al. 2016) and CDRMIP (Keller et al. 2018) experiments that are based on the idealised experiments and have been run with ACCESS-ESM1.5. Selected BGC fields are then discussed for the idealised experiments and their variants.

3.1. Control simulations

The piControl simulation has been run for 500 years with ACCESS-CM2, and 1000 years with ACCESS-ESM1.5. The longer piControl run for ACCESS-ESM1.5 was required for the long reversibility experiment in CDRMIP and the extended length (1000 years) run for abrupt-4xCO2 (r2i1p1f1). Results from the first 300 years of the piControl of both models are shown in Fig. 1.


Fig. 1.  Global physical climate metrics from CMIP idealised experiments. Trends of anomalies in average global surface temperature (top left), net radiative energy flux (top right), sea-ice cover (bottom left) and anomalies in average total ocean temperature (bottom right), from idealised CMIP6 experiments with ACCESS-ESM1.5 (solid lines for transient experiments, black for piControl and green for esm-piControl) and ACCESS-CM2 (dashed lines for transient experiments, red for piControl). 12-month boxcar filters have been applied to all metrics.
Click to zoom

Temperatures are plotted as anomalies relative to the first decade of the respective controls. There is a difference (not shown) of ~0.6 K in the global near-surface air temperature (tas) of the two models at the start of their respective control simulations (years 0950 and 0101 for ACCESS-CM2 and ACCESS-ESM1.5 respectively), with ACCESS-ESM1.5 being warmer; this is likely due to the longer spin-up of ACCESS-ESM1.5. There is a small drift in global tas in both models; a linear fit gives 4.99 × 10−4 K year−1 for ACCESS-CM2 (500 years) and 6.54 × 10−5 K year−1 for ACCESS-ESM1.5 (1000 years). These drifts are close to those of global ocean temperatures (thetaoga, Fig. 1d) of 5.51 × 10−4 K year−1 for ACCESS-CM2 and 6.39 × 10−5 K year−1 for ACCESS-ESM1.5. The drift in total ocean temperature of ACCESS-CM2 is approximately twice the magnitude of an earlier version of ACCESS, ACCESS-1.3, and has the opposite sign (Bi et al. 2020); the drift is substantially less in ACCESS-ESM1.5 due to the long spin-up. Drifts in coupled models are common, and the ACCESS models’ drifts are similar to other CMIP6 models (Irving et al. 2021).

As reported in Ziehn et al. (2020a), the drift in ACCESS-ESM1.5 does not correspond with the TOA net radiation imbalance (Fig. 1b), which shows an energy loss of 0.25 W m−2. Approximately 40% of the global energy imbalance is contributed by CABLE, but the source of the remainder is not known (Ziehn et al. 2020a). Bi et al. (2020) discuss the ACCESS-CM2 TOA radiation imbalance in relation to the total ocean warming, noting that 87% of the ocean’s net gain in energy is due to the TOA imbalance, with the remainder attributed to a lack of energy conservation. These discrepancies in energy conservation are similar to other CMIP6 models (Irving et al. 2021). Fig. 1c also shows global sea-ice area which is relatively stable at ~20 × 1012 m2 for the piControl run from both models, with drifts of less than −0.003 × 1012 m2 year−1.

A second control run, esm-piControl, is required for models with an interactive carbon cycle such as ACCESS-ESM1.5. In this simulation the atmospheric CO2 is not fixed but evolves depending on the net land and ocean carbon fluxes. For a control run, these fluxes should be close to zero (globally and averaged over decades) so that the atmospheric CO2 does not drift. The 500-year esm-piControl run starts after 170 years of spin-up with freely evolving atmospheric CO2, with the spin-up starting from the initial year of the standard piControl. The global mean surface-layer CO2 (Fig. 2) drifts slightly in the esm-piControl run (1 ppm century−1) due to global mean land and ocean carbon fluxes to the atmosphere of −0.03 and 0.05 Pg C year−1 respectively.


Fig. 2.  Global average surface-layer CO2 from emission driven experiments with ACCESS-ESM1.5: esm-piControl, ZECMIP and CDRMIP pulse experiments. For reference, piControl and 1pctCO2 are included. 12-month boxcar filters have been applied to all metrics.
F2

This drift is well within the guidelines for C4MIP of less than 5 ppm century−1 (Jones et al. 2016). The drift in CO2 leads to a larger temperature drift (2.26 × 10−4 K year−1) in the esm-piControl compared to the piControl but this is still very small (as seen in Fig. 1a).

3.2. Climate from idealised experiments

The abrupt-4xCO2 and 1pctCO2 simulations see large increases in atmospheric CO2, producing an increase in temperature, a reduction in sea-ice area and a positive TOA energy balance. Results from these simulations with ACCESS are shown in Fig. 1, with the greater warming in ACCESS-CM2 over ACCESS-ESM1.5 due largely to the different atmospheric configurations. These simulations can be used to calculate a model’s equilibrium climate sensitivity (ECS, based on abrupt-4xCO2; Gregory et al. 2004) and transient climate response (TCR, based on 1pctCO2). For ACCESS-CM2, the ECS and TCR are 4.7 and 2.1°C respectively (Meehl et al. 2020), whereas ACCESS-ESM1.5 has an ECS of 3.9 and TCR of 2.0°C (Ziehn et al. 2020a), which is similar to the climate sensitivities of ACCESS1.0 and ACCESS1.3 (3.6 and 3.9°C respectively; Zelinka et al. 2020) which were submitted to CMIP5.

Both the mean and spread of ECS values among models submitted to CMIP6 are significantly higher than those in CMIP5 (Grose et al. 2020). Furthermore, several CMIP6 models simulate climate responses that do not match ECS estimates from independent assessments using multiple lines of evidence; for example, Sherwood et al. (2020) determined a likely ECS range of 2.3–4.5 K, based on historical and palaeoclimatological records. This overall increase in CMIP6 ECS is largely driven by a subset of models with new or updated physics, such as ACCESS-CM2, which tend to have high ECS values. An investigation of 50 CMIP6 models by Zelinka et al. (2020), in which 16 were found to have an ECS higher than ACCESS-CM2 (including the Met Office models), demonstrated that these ECS increases are primarily due to the representation of cloud microphysics. This provides an explanation as to why ACCESS-CM2 saw the increase but ACCESS-ESM1.5 did not, since the atmospheric components in ACCESS-CM2 have diverged the most since the ACCESS1.3 configuration in CMIP5. Interestingly, the KACE-1-0-G model (Lee et al. 2020) simulated an ECS of 4.5°C (Meehl et al. 2020), very similar to ACCESS-CM2; these two models share the same atmosphere (UM ver. 10.6) and similar ocean model (MOM4.1 in KACE-1-0-G, and MOM5 in ACCESS-CM2). By contrast, HadGEM3-GC31-LL uses the same atmosphere but a different ocean (NEMO ver. 3.6; Storkey et al. 2018), and simulated a significantly higher ECS (ECS = 5.5°C; Senior et al. 2020). This suggests that the ocean and sea-ice components may play a significant role in ECS; however, further investigation is required to understand this complexity.

The larger climate sensitivity of ACCESS-CM2 also shows up in the larger TOA energy imbalance from the increase in atmospheric CO2; this is likely due to a GA7.0 update in the representation of CO2 absorption, resulting in a higher CO2 effective radiative forcing (ERF; discussed further below in relation to RFMIP). As expected, the larger temperature increase in ACCESS-CM2 leads to a larger loss of sea ice, with rapid initial sea-ice loss in both models for the abrupt-4xCO2 case. ACCESS-CM2 shows a further period of rapid loss around years 40–60, driven by the northern polar region and possibly connected to the higher climate sensitivity.

The two idealised ACCESS-CM2 runs (abrupt-4xCO2 and 1pctCO2) reach very similar levels of minimum sea ice, compared to the two idealised ACCESS-ESM1.5 runs where the 1pctCO2 sea-ice area does not drop as low as in the abrupt-4xCO2. This is likely due to the late-simulation increase in surface temperature in the ACCESS-CM2 1pctCO2 run, which outstrips the ACCESS-ESM1.5 1pctCO2 temperature increase after 80 years (see Fig. 1a), likely a result of its higher ECS. Total ocean temperature (Fig. 1d) increases more slowly than the surface air temperature, particularly in the abrupt-4xCO2 simulation, reflecting the time taken for heat to be transported into the deep ocean.

As described in Pincus et al. (2016), a series of atmosphere-only experiments have been performed, as part of RFMIP, to characterise the ERF due to human and natural forcings to the climate. Thirty-year simulations were performed using climatologies of sea surface temperature and sea ice taken from the piControl run and forced with one of the following: pre-industrial conditions (piClim-control), all present-day anthropogenic forcing (piClim-anthro), present-day GHG concentrations (piClim-ghg), present-day aerosols (piClim-aer), present-day land-use forcing separately (piClim-lu; ACCESS-ESM1.5 only) or quadrupled CO2 (piClim-4xCO2; which is subsequently scaled by a factor of 0.2266 to be equivalent to the present-day CO2 concentration of 1.4× pre-industrial CO2, and is used to separate out the forcing from CO2 from other greenhouse gases). The ERF values from these experiments, and the scaled 1.4xCO2 value, are presented in Table 7. ACCESS-CM2 simulated a significantly higher 4xCO2 ERF than ACCESS-ESM1.5, and very close to the CMIP6 mean of 7.98 ± 0.38 W m−2 (Smith et al. 2020). This is mostly due to an improvement in the representation of CO2 absorption in the band 8–13 μm, which was implemented in GA7.0 and adopted into ACCESS-CM2 (Walters et al. 2019); this is also likely a contributing factor to the high ECS of ACCESS-CM2. ACCESS-ESM1.5 simulated a 4xCO2 ERF of 7.04 W m−2, similar to that of the CMIP5 models HadGEM2-ES (6.99 W m−2, which uses a similar atmosphere; Andrews et al. 2012) and ACCESS1.3 (6.75 W m−2, the predecessor of ACCESS-ESM1.5). The GHG, aerosol, athropogenic and land-use (ACCESS-ESM1.5 only) ERF values for both ACCESS models are similar to those of HadGEM3-GC31-LL, and sit well within the spread of CMIP6 values (Smith et al. 2020).


Table 7.  Effective radiative forcing from RFMIP experiments (W m−2).
Click to zoom

ACCESS-ESM1.5 has been used to run experiments for C4MIP (including ZECMIP) and CDRMIP that are variants of the 1ptCO2 simulation, or branch from the 1pctCO2 simulation; many of these utilise the interactive carbon cycle. Global average surface CO2 concentration for emissions-driven experiments are shown in Fig. 2 and global mean tas for all C4MIP and CDRMIP experiments is shown in Fig. 3. The response of the carbon cycle is discussed in Section 2.3. For climate–carbon feedback analysis, 1pctCO2 simulations are performed where only the radiation scheme (1pctCO2-rad) or only the biogeochemistry (1pctCO2-bgc) is subject to increasing atmospheric CO2. As expected, tas increases in the 1pctCO2-rad experiment (only slightly less than in 1pctCO2) whereas the 1pctCO2-bgc simulation shows only a very small temperature increase. The small differences in temperature response between 1pctCO2-rad and 1pctCO2 and between 1pctCO2-bgc and piControl are due to feedbacks between the carbon cycle and climate, principally through changes in the evolution of leaf area index and surface albedo. The ACCESS-ESM1.5 temperature response is in the middle of the range of CMIP6 models (Arora et al. 2020).


Fig. 3.  Trends in anomalies of average global surface temperature from extra idealised experiments with ACCESS-ESM1.5. For clarity, emission driven experiments (esm-piControl, esm-pi-CO2pulse and esm-pi-cdr-pulse) start at year 0201 rather than 0001. 12-month boxcar filters have been applied to all metrics.
F3

The C4MIP-ZECMIP experiments (Jones et al. 2019) explore the response of the climate to an abrupt shift to zero CO2 emissions. Simulations start at various points from 1pctCO2, a prescribed concentration experiment, and continue as emissions-driven experiments (with zero emissions). The temperature response is small in all cases, with a slight cooling for the lowest branch, almost no change in temperature for the middle branch and a small warming for the highest branch. Most CMIP6 models show more cooling than ACCESS-ESM1.5 but the spread across models (as given by the standard deviation) is larger than the mean or median cooling (MacDougall et al. 2020).

ACCESS-ESM1.5 has completed three CDRMIP experiments. The reversibility of the climate system is tested in 1pctCO2-cdr by extending the 1pctCO2 simulation for 140 years with atmospheric CO2 reducing by 1% per year, and then running for 620 years with pre-industrial CO2 (noting only 220 years are shown in Fig. 3). With similar results to Ziehn et al. (2020d), in which the previous ACCESS-ESM version was used, global mean tas remains ~1 K above the pre-industrial temperature at the end of the CO2 decrease and only slowly decreases over the following centuries, remaining 0.3 K above pre-industrial after 600 years.

A further two CDRMIP experiments are perturbations to the esm-piControl, and are run in emissions-driven mode. Atmospheric CO2 is instantaneously increased or decreased by 47 ppm (equivalent to 100 Pg C) to examine the response to the perturbation over 90 years and whether that response is symmetric. The magnitudes of differences in global mean tas are small (<±0.4 K, see Fig. 3), with the positive temperature difference (due to the positive CO2 perturbation) being somewhat larger and shorter lived than the negative temperature difference from the negative CO2 perturbation. However, another positive temperature difference occurs in the last decade of the simulation due to variability and a decrease in temperature in the esm-piControl. This suggests that an ensemble (currently planned) is needed to better constrain the temperature response.

3.3. Biogeochemistry in idealised experiments

Global mean ACCESS-ESM1.5 biogeochemical metrics from CMIP6 idealised experiments are shown in Fig. 4 and 5. Fig. 4 includes DECK experiments (piControl, abrupt-4xCO2 and 1pctCO2) as well as experiments where the atmospheric CO forcing in the 1pctCO2 simulation is applied separately to the radiation or biogeochemical components of the model. Fig. 5 shows experiments that branch from 1pctCO2 (including experiments associated with ZECMIP and CDRMIP) as well as emission-driven experiments (esm-piControl, pulse experiments).


Fig. 4.  Global biogeochemical metrics from idealised experiments: the piControl, abrupt-4xCO2, 1pctCO2, 1pctCO2-bgc and 1pctCO2-rad. Trends are shown of carbon flux into the land (top left), net land primary productivity (top right), carbon flux into the ocean (middle left), primary ocean productivity (middle right), oxygen flux into the ocean (bottom left) and average sea surface pH (bottom right) from idealised CMIP experiments with ACCESS-ESM1.5. Sixty-month boxcar filters have been applied to remove seasonal cycles and reduce interannual variability in all metrics, except ocean carbon flux and pH which were filtered with a 12-month boxcar to better reveal the differences between simulations for these variables.
Click to zoom


Fig. 5.  Global biogeochemical metrics from idealised experiments; the piControl, esm-piControl, 1pctCO2-rev, zero-emission and pulse experiments, with the same layout as Fig. 4. For clarity, emission driven experiments (esm-piControl, pulse and pulse-cdr) start at year 0201 rather than 0001.
Click to zoom

Both Fig. 4 and 5 show CO2 flux and productivity from the land (top rows) and the ocean (middle); ocean O2 flux and acidification are in the bottom rows. Metrics are filtered by 5- or 1-year box-car filters to remove the seasonal cycle and to reduce large interannual variability, in order to better reveal the differences between simulations.

3.3.1. Land biogeochemistry

The net land carbon flux (Fig. 4a) is the difference between photosynthesis and respiration, and is generally small (in the time mean) except where atmospheric CO2 is changing. The land net primary productivity (NPP, Fig. 4b) shows carbon uptake by plants (which is approximately balanced by soil respiration). The abrupt-4xCO2 simulation shows uptake of carbon to the land initially, and a large increase in NPP because photosynthesis responds more quickly to the increased CO2 than does respiration. Although the net land carbon flux returns close to zero within ~40 years, NPP (Fig. 4b) remains ~18% higher than the control due to the elevated atmospheric CO2. This indicates that soil respiration is also elevated relative to the control, given the approximately zero net land flux. Increasing CO2 in the 1pctCO2 experiment produces increasing NPP through most of the experiment, though the rate of increase declines after ~30 years. This leads to a peak in the net land carbon flux after 20–30 years and then a decline back to around zero mean flux after 100 years. This return to zero land flux is unusual compared to most other CMIP6 Earth System Models (Arora et al. 2020), which tend to produce net land carbon uptake throughout the simulation; the multimodel mean land carbon flux from fig. 2 of Arora et al. (2020) is 3.8 Pg C year−1 at year 0140 (when the 1pctCO2 atmospheric CO2 reaches four times pre-industrial CO2). Across these 140 years, the cumulative land carbon uptake for ACCESS-ESM1.5 (215 Pg C) is lower than other models (408–1204 Pg C); models with nitrogen limitation give lower carbon uptake (mean 536 Pg C) compared with those without (mean 754 Pg C) (Arora et al. 2020, fig. 4). ACCESS-ESM1.5 has been run with both nitrogen and phosphorus limitation, and Ziehn et al. (2021b) shows that this contributes to the smaller land carbon uptake.

The separate impacts of warming and increased CO2 on the land carbon fluxes can be seen in the 1pctCO2-rad and 1pctCO2-bgc experiments. After an initial rise of 10 Pg C year−1 over the first 30 years, NPP increases (if slowly) throughout the 1pctCO2-bgc experiment due to increasing CO2 but no significant change in temperature (Fig. 3). This is in contrast to the 1pctCO2 run in which NPP declines slightly after approximately 120 years. The impact of warming is confirmed by the 1pctCO2-rad experiment, where NPP decreases substantially when warming occurs but the carbon cycle does not experience increasing CO2. The response in NPP is also reflected in the response of the net land carbon flux. After the initial increase in land carbon uptake, the decrease in net land carbon uptake in the 1pctCO2-bgc case is slower than in the 1pctCO2 case, remaining ~1 Pg C year−1 at the end of the simulation. The 1pctCO2-rad experiment shows an increasing carbon release from the land as expected. While following the general response of other CMIP6 models in these experiments, ACCESS-ESM1.5 experiences less carbon uptake than other models in the 1pctCO2-bgc experiment, and releases more carbon than almost all models in the 1pctCO2-rad experiment (Arora et al. 2020, fig. A1).

The carbon fluxes for the four experiments that branch from the 1pctCO2 simulation (three ZECMIP and CDRMIP reversibility) are shown in Fig. 5. The ZECMIP experiments show a very small decline in land NPP relative to the 1pctCO2 simulation, but this decline is not seen in the net land flux. Given the relatively small changes in temperature and atmospheric CO2 simulated in each of these branch experiments (Fig. 2, 3), it is to be expected that a land biosphere approximately in balance (as indicated by the near zero net land flux in 1pctCO2) would remain in balance and continue to give approximately zero net land flux.

Where the ZECMIP experiments show the response of the carbon fluxes to a relatively slow decrease in atmospheric CO2, the CDRMIP reversibility (1pctCO2-cdr) experiment shows the response to a relatively large decrease. The net land flux initially remains close to zero, but from around year 0220 becomes increasingly negative, indicating a source of carbon to the atmosphere. The maximum source occurs around year 0280, coincident with atmospheric CO2 returning to pre-industrial levels, after which the flux returns to approximately zero over ~30 years. This behaviour is reflected in the land NPP, which initially decreases slowly, then more rapidly, dropping below the pre-industrial NPP around year 0280, before returning close to, but lower than, pre-industrial NPP. The overshoot and return to pre-industrial land fluxes likely results from the slower return of temperature to pre-industrial compared with atmospheric CO2. The overshoot to below pre-industrial net land carbon flux and NPP is widespread globally but analysis of an earlier, similar simulation by Ziehn et al. (2020d) showed a variation in response timing at regional scale. Responses varied due to the dominant vegetation type in a region, whether a region was driven more strongly by variations in temperature or precipitation and from the interaction of carbon pools with different turnover times. These regional responses are being further explored in a multimodel study in preparation.

Also shown in Fig. 5 are the BGC responses to the CDRMIP pulse experiments. The abrupt increase or decrease in atmospheric CO2 (Fig. 2) leads to positive or negative pulses of land CO2 fluxes. When CO2 is added to the atmosphere, the NPP increases by ~20%, decreasing to ~5% above pre-pulse levels by the end of the simulation. The land takes up 8 Pg C in the year that the pulse of CO2 is put into the atmosphere, but the large uptake is short-lived. When CO2 is removed, the decrease in NPP is 15–20% over 6 years before increasing to ~3% below pre-pulse levels. The loss of carbon from the land is 5–6 Pg C for the first 5 years after the negative pulse of atmospheric CO2. Of the 100 Pg C added to the atmospheric CO2, ~40 Pg C is taken up by the land with most of the uptake occurring in the first 40 years. When 100 Pg C is removed from the atmosphere, the land loses ~50 Pg C over 40 years but then slowly takes up carbon, giving a loss of ~40 Pg C by the end of the simulation. The relatively small asymmetry in global land carbon response from the positive and negative CO2 pulses should be confirmed regionally, and the analysis of these simulations would benefit from an ensemble of runs, which are currently planned.

3.3.2. Ocean biogeochemistry

The latter four panels in Fig. 4 and 5 capture various aspects of the ocean biogeochemistry in the idealised experiments run with ACCESS-ESM1.5. The surface carbon flux into the ocean (Fig. 4c, 5c) is driven by the difference in CO2 partial pressure across the ocean–atmosphere surface, and is largely controlled by physical processes and the changing atmospheric CO2. Ocean productivity (Fig. 4d, 5d) is mainly constrained by the supply of nutrients into the upper ocean, where solar radiation can drive biological growth. The O2 flux across the ocean surface (Fig. 4e, 5e) is also related to ocean circulation, as well as temperature. Trends in surface acidity (Fig. 4f, 5f) are dominated by the atmospheric CO2, with some influence from the existing alkalinity of the surface ocean. The CO2 flux and ocean acidification show the least variability, being almost a direct response to atmospheric forcing. Other metrics (e.g. productivity and O2 flux) are more sensitive to other conditions, such as temperature and ocean dynamics, and are more variable.

The largest carbon fluxes into the ocean are in the abrupt-4xCO2 experiment, which imposes a high constant atmospheric CO2 boundary condition and results in a flux into the ocean that reduces after the initial shock. A high CO2 flux also occurs in the 1pctCO2 experiment, where the increasing atmospheric CO2 drives an increasing flux over the first few decades. A maximum flux in 1pctCO2 is reached when the value of the Revelle factor increases (e.g. Jiang et al. 2019), reducing the efficiency of CO2 uptake into the ocean. Mean ocean uptake over the last 30 years of the simulation is 5.3 Pg C year−1 which is slightly larger than the CMIP6 multimodel mean of 5.0 Pg C year−1 (Arora et al. 2020, fig. 2).

In experiments that branch from 1pctCO2 (i.e. ZECMIP and 1pctCO2-cdr), the ocean takes up CO2 relatively strongly at the branch points (Fig. 5c). In the ZECMIP experiments this uptake decreases after the change to zero emissions, initially rapidly and then more slowly. This is driven by the stabilisation of the atmospheric CO2 and the reduction in the difference between ocean and atmospheric CO2 partial pressures. In 1pctCO2-cdr, the ocean changes from a sink to a source after 30 years because the atmospheric CO2 drops below the average partial pressure of CO2 in the surface ocean. The source increases in magnitude until atmospheric CO2 stabilises at pre-industrial levels, after which the ocean flux slowly returns towards zero.

Also shown in Fig. 5 are the BGC responses to the CDRMIP pulse experiments. The abrupt increase or decrease in atmospheric CO2 (Fig. 2) leads to positive or negative pulses of ocean carbon fluxes, similar to those seen in the land carbon fluxes. The response is reasonably symmetric for the ocean flux, with the pulse decaying over ~20 years. Of the 100 Pg C added to or removed from the atmosphere, the ocean carbon absorption or release is ~30 Pg C by the end of the run, with 50% of the total absorption or release occurring in the first 5 years.

Average ocean productivity is 33.5 Pg C year−1 in the piControl. Initial reductions in productivity of the order of ~2 Pg C year−1 are evident in both the abrupt-4xCO2 and 1pctCO2 experiments (Fig. 4d). Interestingly, these two experiments show an increase in productivity after their initial decrease and, by the end of the abrupt-4xCO2 case, global productivity becomes greater than that of the piControl experiment. There is considerable uncertainty in the productivity response from high-end climate change scenarios; the standard deviation between CMIP6 models of changes in global ocean productivity in ssp585 experiments is greater than the multimodel mean (Kwiatkowski et al. 2020). Preliminary analysis shows that ocean productivity changes are not a simple, uniform global response, but rather that local productivity can increase or decrease in different regions driven by regional changes in temperature, mixed layer depth or nutrient supply (not shown in the global analysis).

In 1pctCO2-bgc, the CO2 fertilisation experiment, there is negligible impact on productivity and O2 fluxes compared with 1pctCO2, whereas in 1pctCO2-rad productivity and O2 flux respond in the same way as 1pctCO2. There is no CO2 fertilisation process in the ocean biogeochemistry component of the ACCESS-ESM1.5. Hence, productivity in 1pctCO2-bgc follows the piControl, and in 1pctCO2-rad follows 1pctCO2.

The O2 flux trends (Fig. 4e, 5e) respond to changes in ocean circulation, temperature and productivity but, unlike CO2 fluxes, there is no change in the direct atmospheric forcing for O2. There is a non-zero O2 flux in piControl (~−76.4 Tmol O2 year−1) due to the remineralisation of detritus in regions with zero O2 in the model, as discussed in Ziehn et al. (2020a). In abrupt-4xCO2, the balance in the O2 flux is broken by rapid stratification of the ocean, increasing the vertical stability in the ocean and shutting down the uptake of O2. There is an increase in O2 outgassing in 1pctCO2 and 1pctCO2-rad as both stratification and surface temperature rise, which is equivalent to the deoxygenation of the ocean in climate change scenarios of ACCESS and other CMIP6 models discussed by Kwiatkowski et al. (2020). The O2 outgassing reduces in all experiments branching from 1pctCO2 (Fig. 5e). In 1pctCO2-cdr, O2 flux reaches parity with piControl after ~100 years, followed by a relative influx for ~50 years until atmospheric CO2 is reduced to pre-industrial levels. This late influx of O2 is small relative to the outgassing, such that the total ocean O2 content is reduced by ~5% relative to the pre-industrial content.

Ocean acidity is not a direct model output of ACCESS-ESM1.5, but can be diagnosed from other ocean carbon variables; the metric previously has been used in model intercomparisons of stresses on marine ecosystems (e.g. Bopp et al. 2013; Kwiatkowski et al. 2020). Acidity is calculated from dissolved inorganic carbon, alkalinity and equilibrium solubility constants. These constants are functions of temperature and salinity, and are determined with the same subroutines used to calculate air–sea CO2 fluxes within the ocean model. Trends in surface ocean acidity (Fig. 4f, 5f) closely follow the atmospheric CO2 boundary condition (Fig. 2), with minimal variability. Notable from the trends shown, and particularly for the branching experiments (ZECMIP and CDRMIP-pulse experiments), is that ocean acidification and atmospheric CO2 are slow to recover from their perturbed states, certainly beyond the 100-year span of the experiments shown. Trends in surface alkalinity due to changes in ocean structure and circulation have a small influence on average acidity as seen in the difference between 1pctCO2-rad and piControl, and that acidity in 1pctCO2-cdr does not return to pre-industrial values.


4. Conclusions

The CMIP6 submission of ACCESS simulation data is much more extensive than the CMIP5 submission, which utilised the models ACCESS1.0 and ACCESS1.3 (Dix et al. 2013). A number of factors contributed to this increase, including the greater scope of CMIP6, the additional capability of the models (ACCESS-ESM1.5 is the first Australian submission with carbon cycle capability), and sufficient compute and data storage resources at the modelling group’s disposal. Compute resources were larger than originally planned due to the awarding of an Australasian Leadership Computing Grant, which enabled DAMIP simulations and an increase in ensemble size for historical and ScenarioMIP experiments. The CMIP6 submission is a major achievement for a relatively small core team especially since the models used for CMIP6 differed more from each other than the two model versions used for CMIP5 and hence had different requirements for their experimental set-up. The OMIP submission would not have been achieved without the COSIMA contribution.

While the ACCESS-ESM1.5 submission was primarily undertaken for the purpose of including Earth system components, the different atmospheric configuration and consequently lower climate sensitivity (compared to ACCESS-CM2) provides good opportunities for comparative studies of the climate generated by each version. The lower compute resource requirement for ACCESS-ESM1.5 has also facilitated additional model use, such as the university-led palaeoclimate simulations contributed to PMIP (Yeung et al. 2021; Choudhury et al. 2022).

It is difficult to fully track the uptake and use of CMIP6 datasets. The Earth System Grid Federation provides statistics of dataset publications and downloads,20 noting that in this context each variable for each experiment is counted as a separate dataset, with further versions also counting as additional datasets. These statistics show that there have been over 51 million downloads of the ACCESS datasets (accessed 11 April 2022) since the first datasets were submitted in December 2019. These download numbers are comparable to models from major international modelling groups, with ACCESS-ESM1.5 in the top 10 models using this download metric.

The number of publications using ACCESS CMIP6 datasets is increasing. The ACCESS-CM2 AMIP simulations are the basis of a paper by Bodman et al. (2020), and an evaluation of historical climate variability and change, as simulated by the ACCESS models, is presented in Rashid et al. (2022). Ziehn et al. (2021b) explores land carbon-concentration and carbon–climate feedbacks using ACCESS-ESM1.5 simulations, whereas Holmes et al. (2022) provides an analysis of ocean warming in ACCESS-CM2. ACCESS data have also been used in many multimodel analyses (e.g. Watterson et al. 2021; Huang et al. 2022). CMIP6 publications are tracked at the CMIP Publication Hub,21 which lists 208 publications (as of 8 April 2022), though tracking is dependent on authors adding their publication details, optionally with the models that have been included in their analysis. The number of publications that list the inclusion of ACCESS-CM2 or ACCESS-ESM1.5 is comparable to the number listed for other well-respected models. Papers that have used ACCESS-OM2 code or data are also tracked independently,22 indicating 35 such publications.

ACCESS-CMIP6 data are also now being used as forcing inputs for regional downscaling activities, including CORDEX-CMIP623 (Gutowski et al. 2016), COWCLIP24 and Australian national and state climate projections (e.g. Di Virgilio et al. 2022). Further use of the ACCESS models and their CMIP6 datasets is welcomed, with information on the availability of ACCESS software and data provided in the next section.


5. Code and data availability

Owing to licensing restrictions, we cannot provide either the source code or documentation papers for the UM, which is used under a license and collaborative partnership. For information on obtaining a license for the UM go to http://www.metoffice.gov.uk/research/collaboration/um-collaboration. CABLE is distributed under an open source license. It is hosted at NCI, and registration is required to access the code repository. Details can be found at https://trac.nci.org.au/trac/cable/wiki/CABLE_Registration. MOM5 is available to download from https://github.com/mom-ocean/MOM5 (ACCESS-CM2) and https://github.com/COSIMA/ACCESS-ESM1.5-MOM5 (ACCESS-ESM1.5, with WOMBAT). CICE is available at https://github.com/COSIMA/cice5 (ACCESS-CM2, ACCESS-OM2) and https://github.com/COSIMA/cice4 (ACCESS-ESM1.5). The OASIS3-MCT coupler is available at https://github.com/COSIMA/oasis3-mct. ACCESS-OM2 code is available at https://github.com/COSIMA/access-om2 and the version at the time of OMIP submission is available at https://github.com/COSIMA/access-om2/releases/tag/2020-11-12. The ACCESS-OM2 configurations used for CMIP6 submission include the following: https://github.com/COSIMA/1deg_jra55_iaf/tree/omip2 and https://github.com/COSIMA/1deg_jra55_iaf/tree/omip2spunup for ACCESS-OM2; and https://github.com/COSIMA/025deg_jra55_iaf/tree/omip_amoctopo_cycle1 through to https://github.com/COSIMA/025deg_jra55_iaf/tree/omip_amoctopo_cycle6 for ACCESS-OM2-025. Rose/cylc suits, which each contain the experiment configuration of a single ACCESS-CM2 realisation, are held in a revision-controlled repository service hosted at the UKMO. A mapping between suite names and CMIP6 ensemble members can be found at https://confluence.csiro.au/display/ACCESS/CMIP6+Archive+-+ACCESS-CM2. Payu configurations for the ACCESS-ESM1.5 experiments piControl, historical and PMIP can be found at the CLEX GitHub repository https://github.com/coecms (e.g. https://github.com/coecms/esm-lig). Configurations for ACCESS-OM2 can be found at the COSIMA GitHub repository https://github.com/COSIMA. The APP4 post-processing code used to CMORise ACCESS model output is available at https://git.nci.org.au/cm2704/APP4 (Mackallah et al. 2022). The mapping between ACCESS output fields and CMIP6 variables is specified in the file APP4/input_files/master_map[_om2].csv. Both python2 and cdms2 will soon be no longer supported, therefore reproducibility is enabled and ensured through a Conda environment. Future development and access to this tools will be facilitated by the ACCESS-NRI.25 CMORised simulation data are published on the Earth System Grid at Dix et al. (2019a) (ACCESS-CM2), Ziehn et al. (2019c) (ACCESS-ESM1.5), Hayashida et al. (2021) (ACCESS-OM2) and Holmes et al. (2021) (ACCESS-OM2-025). See also https://esgf.nci.org.au/search/cmip6-nci (ESGF data portal, NCI node) and https://doi.org/10.25914/5e6acd0492b39 (NCI GeoNetwork record). NCI users can also access the data directly from project fs38 (https://my.nci.org.au/mancini/project/fs38). Model output, the semi-processed simulation data which include many variables not requested by CMIP6, that has not been CMORised is also available for NCI users in project p73 (https://my.nci.org.au/mancini/project/p73).


Data availability

CMORised simulation data are published on the Earth System Grid at Dix et al. (2019a) (ACCESS-CM2), Ziehn et al. (2019c) (ACCESS-ESM1.5), Hayashida et al. (2021) (ACCESS-OM2) and Holmes et al. (2021) (ACCESS-OM2-025). See also https://esgf.nci.org.au/search/cmip6-nci (ESGF data portal, NCI node) and https://doi.org/10.25914/5e6acd0492b39 (NCI GeoNetwork record). NCI users can also access the data directly from project fs38 (https://my.nci.org.au/mancini/project/fs38). Model output, the semi-processed simulation data which include many variables not requested by CMIP6, that has not been CMORised is also available for NCI users in project p73 (https://my.nci.org.au/mancini/project/p73).


Conflicts of interest

Josephine Brown is an Associate Editor for the Journal of Southern Hemisphere Earth Systems Science but did not at any stage have editor-level access to this manuscript while in peer review, as is the standard practice when handling manuscripts submitted by an editor to this journal. The authors have no further conflicts of interest to declare.


Declaration of funding

ACCESS simulations undertaken with the assistance an Australasian Leadership Computing Grant by sponsor, the National Computational Infrastructure (NCI Australia) (grant #12840 under ALCG2020, NCI project gy57). NCI Australia had a governance role in the publication of ACCESS data to CMIP6, and NCI staff contributed to the preparation of this manuscript. COSIMA and A. E. Kiss were funded by Australian Research Council grants LP160100073 and LP200100406. A. E. Kiss was also funded by the Australian Government's Australian Antarctic Science Program grant 4541. This research was undertaken with the assistance of resources from the NCI, which is supported by the Australian Government.



Acknowledgements

We acknowledge the World Climate Research Programme, which, through its Working Group on Coupled Modelling, coordinated and promoted CMIP6. We thank the Earth System Grid Federation (ESGF) for archiving the data and providing access, and the multiple funding agencies who support CMIP6 and ESGF. The ACCESS CMIP6 submission work was jointly funded through CSIRO and the Earth Systems and Climate Change Hub and subsequent Climate Systems Hub of the Australian Government’s National Environmental Science Program (NESP). ACCESS simulations, data processing and data publication were undertaken with the assistance of resources from NCI Australia, a NCRIS-enabled (National Collaborative Research Infrastructure Strategy) capability supported by the Australian Government. We also thank Rui Yang and Paul Leopardi at NCI Australia for their work to enable ACCESS modelling on Raijin and Gadi. The UM software is used under the auspices of a partnership agreement between the UK Met Office, CSIRO and the Australian Bureau of Meteorology. This work was built on the allied research efforts of the host institutions, the ARC Centre of Excellence for Climate Extremes (CLEX) and the Consortium for Ocean–Sea-Ice Modelling in Australia (COSIMA).


References

Andrews T, Gregory JM, Webb MJ, Taylor KE (2012) Forcing, feedbacks and climate sensitivity in CMIP5 coupled atmosphere–ocean climate models. Geophysical Research Letters 39, L09712
Forcing, feedbacks and climate sensitivity in CMIP5 coupled atmosphere–ocean climate models.Crossref | GoogleScholarGoogle Scholar |

Arfeuille F, Weisenstein D, Mack H, Rozanov E, Peter T, Brönnimann S (2014) Volcanic forcing for climate modeling: a new microphysics-based data set covering years 1600–present. Climate of the Past 10, 359–375.
Volcanic forcing for climate modeling: a new microphysics-based data set covering years 1600–present.Crossref | GoogleScholarGoogle Scholar |

Arora VK, Katavouta A, Williams RG, Jones CD, Brovkin V, Friedlingstein P, Schwinger J, Bopp L, Boucher O, Cadule P, Chamberlain MA, Christian JR, Delire C, Fisher RA, Hajima T, Ilyina T, Joetzjer E, Kawamiya M, Koven CD, Krasting JP, Law RM, Lawrence DM, Lenton A, Lindsay K, Pongratz J, Raddatz T, Séférian R, Tachiiri K, Tjiputra JF, Wiltshire A, Wu T, Ziehn T (2020) Carbon-concentration and carbon–climate feedbacks in CMIP6 models and their comparison to CMIP5 models. Biogeosciences 17, 4173–4222.
Carbon-concentration and carbon–climate feedbacks in CMIP6 models and their comparison to CMIP5 models.Crossref | GoogleScholarGoogle Scholar |

Balaji V, Maisonnave E, Zadeh N, Lawrence BN, Biercamp J, Fladrich U, Aloisio G, Benson R, Caubel A, Durachta J, Foujols M-A, Lister G, Mocavero S, Underwood S, Wright G (2017) CPMIP: measurements of real computational performance of Earth system models in CMIP6. Geoscientific Model Development 10, 19–34.
CPMIP: measurements of real computational performance of Earth system models in CMIP6.Crossref | GoogleScholarGoogle Scholar |

Balaji V, Taylor KE, Juckes M, Lawrence BN, Durack PJ, Lautenschlager M, Blanton C, Cinquini L, Denvil S, Elkington M, Guglielmo F, Guilyardi E, Hassell D, Kharin S, Kindermann S, Nikonov S, Radhakrishnan A, Stockhause M, Weigel T, Williams D (2018) Requirements for a global data infrastructure in support of CMIP6. Geoscientific Model Development 11, 3659–3680.
Requirements for a global data infrastructure in support of CMIP6.Crossref | GoogleScholarGoogle Scholar |

Bellouin N, Collins W, Culverwell I, Halloran P, Hardiman S, Hinton T, Jones C, McDonald R, McLaren A, O’Connor F, et al. (2011a) The HadGEM2 family of met office unified model climate configurations. Geoscientific Model Development 4, 723

Bellouin N, Rae J, Jones A, Johnson C, Haywood J, Boucher O (2011b) Aerosol forcing in the Climate Model Intercomparison Project (CMIP5) simulations by HadGEM2-ES and the role of ammonium nitrate. Journal of Geophysical Research: Atmospheres 116, D20206
Aerosol forcing in the Climate Model Intercomparison Project (CMIP5) simulations by HadGEM2-ES and the role of ammonium nitrate.Crossref | GoogleScholarGoogle Scholar |

Berger A, Loutre MF (1991) Insolation values for the climate of the last 10 million years. Quaternary Science Reviews 10, 297–317.
Insolation values for the climate of the last 10 million years.Crossref | GoogleScholarGoogle Scholar |

Best MJ, Pryor M, Clark DB, Rooney GG, Essery RLH, Ménard CB, Edwards JM, Hendry MA, Porson A, Gedney N, et al. (2011) The Joint UK Land Environment Simulator (JULES), model description – part 1: energy and water fluxes. Geoscientific Model Development 4, 677–699.
The Joint UK Land Environment Simulator (JULES), model description – part 1: energy and water fluxes.Crossref | GoogleScholarGoogle Scholar |

Bi D, Dix M, Marsland SJ, O’Farrell S, Rashid HA, Uotila P, Hirst AC, Kowalczyk E, Golebiewski M, Sullivan A, Yan H, Hannah N, Franklin C, Sun Z, Vohralik P, Watterson I, Zhou X, Fiedler R, Collier M, Ma Y, Noonan J, Stevens L, Uhe P, Zhu H, Griffies SM, Hill R, Harris C, Puri K (2013) The ACCESS coupled model: description, control climate and evaluation. Australian Meteorological and Oceanographic Journal 63, 41–64.
The ACCESS coupled model: description, control climate and evaluation.Crossref | GoogleScholarGoogle Scholar |

Bi D, Dix M, Marsland S, O’Farrell S, Sullivan A, Bodman R, Law R, Harman I, Srbinovsky J, Rashid HA, Dobrohotoff P, Mackallah C, Yan H, Hirst A, Savita A, Dias FB, Woodhouse M, Fiedler R, Heerdegen A (2020) Configuration and spin-up of ACCESS-CM2, the new generation Australian Community Climate and Earth System Simulator Coupled Model. Journal of Southern Hemisphere Earth Systems Science 70, 225–251.
Configuration and spin-up of ACCESS-CM2, the new generation Australian Community Climate and Earth System Simulator Coupled Model.Crossref | GoogleScholarGoogle Scholar |

Bodman RW, Karoly DJ, Dix MR, Harman IN, Srbinovsky J, Dobrohotoff PB, Mackallah C (2020) Evaluation of CMIP6 AMIP climate simulations with the ACCESS-AM2 model. Journal of Southern Hemisphere Earth Systems Science 70, 166–179.
Evaluation of CMIP6 AMIP climate simulations with the ACCESS-AM2 model.Crossref | GoogleScholarGoogle Scholar |

Bopp L, Resplandy L, Orr JC, Doney SC, Dunne JP, Gehlen M, Halloran P, Heinze C, Ilyina T, Séférian R, Tjiputra J, Vichi M (2013) Multiple stressors of ocean ecosystems in the 21st century: projections with CMIP5 models. Biogeosciences 10, 6225–6245.
Multiple stressors of ocean ecosystems in the 21st century: projections with CMIP5 models.Crossref | GoogleScholarGoogle Scholar |

Checa-Garcia R, Hegglin MI, Kinnison D, Plummer DA, Shine KP (2018) Historical tropospheric and stratospheric ozone radiative forcing using the CMIP6 database. Geophysical Research Letters 45, 3264–3273.
Historical tropospheric and stratospheric ozone radiative forcing using the CMIP6 database.Crossref | GoogleScholarGoogle Scholar |

Choudhury D, Menviel L, Meissner KJ, Yeung NKH, Chamberlain M, Ziehn T (2022) Marine carbon cycle response to a warmer Southern Ocean: the case of the last interglacial. Climate of the Past 18, 507–523.
Marine carbon cycle response to a warmer Southern Ocean: the case of the last interglacial.Crossref | GoogleScholarGoogle Scholar |

Clark DB, Mercado LM, Sitch S, Jones CD, Gedney N, Best MJ, Pryor M, Rooney GG, Essery RLH, Blyth E, et al. (2011) The Joint UK Land Environment Simulator (JULES), model description – part 2: carbon fluxes and vegetation dynamics. Geoscientific Model Development 4, 701–722.
The Joint UK Land Environment Simulator (JULES), model description – part 2: carbon fluxes and vegetation dynamics.Crossref | GoogleScholarGoogle Scholar |

Collier M, Uhe P (2012) CMIP5 datasets from the ACCESS1.0 and ACCESS1.3 coupled climate models. CAWCR Technical Report Number 059. Available at https://www.cawcr.gov.au/technical-reports/CTR_059.pdf

Deser C, Lehner F, Rodgers KB, Ault T, Delworth TL, DiNezio PN, Fiore A, Frankignoul C, Fyfe JC, Horton DE, et al. (2020) Insights from Earth system model initial-condition large ensembles and future prospects. Nature Climate Change 10, 277–286.
Insights from Earth system model initial-condition large ensembles and future prospects.Crossref | GoogleScholarGoogle Scholar |

Di Virgilio G, Ji F, Tam E, Nishant N, Evans JP, Thomas C, Riley ML, Beyer K, Grose MR, Narsey S, Delage F (2022) Selecting CMIP6 GCMs for CORDEX dynamical downscaling: model performance, independence, and climate change signals. Earth’s Future 10, e2021EF002625
Selecting CMIP6 GCMs for CORDEX dynamical downscaling: model performance, independence, and climate change signals.Crossref | GoogleScholarGoogle Scholar |

Dix M, Vohralik P, Bi D, Rashid H, Marsland S, O’Farrell S, Uotila P, Hirst T, Kowalczyk E, Sullivan A, Yan H, Franklin C, Sun Z, Watterson I, Collier M, Noonan J, Rotstayn L, Stevens L, Uhe P, Puri K (2013) The ACCESS coupled model: documentation of core CMIP5 simulations and initial results. Australian Meteorological and Oceanographic Journal 63, 83–99.
The ACCESS coupled model: documentation of core CMIP5 simulations and initial results.Crossref | GoogleScholarGoogle Scholar |

Dix M, Bi D, Dobrohotoff P, Fiedler R, Harman I, Law R, Mackallah C, Marsland S, O’Farrell S, Rashid H, Srbinovsky J, Sullivan A, Trenham C, Vohralik P, Watterson I, Williams G, Woodhouse M, Bodman R, Dias FB, Domingues C, Hannah N, Heerdegen A, Savita A, Wales S, Allen C, Druken K, Evans B, Richards C, Ridzwan SM, Roberts D, Smillie J, Snow K, Ward M, Yang R (2019a) CSIRO-ARCCSS ACCESS-CM2 model output prepared for CMIP6 CMIP.
| Crossref |

Dix M, Bi D, Dobrohotoff P, Fiedler R, Harman I, Law R, Mackallah C, Marsland S, O’Farrell S, Rashid H, Srbinovsky J, Sullivan A, Trenham C, Vohralik P, Watterson I, Williams G, Woodhouse M, Bodman R, Dias FB, Domingues C, Hannah N, Heerdegen A, Savita A, Wales S, Allen C, Druken K, Evans B, Richards C, Ridzwan SM, Roberts D, Smillie J, Snow K, Ward M, Yang R (2019b) CSIRO-ARCCSS ACCESS-CM2 model output prepared for CMIP6 ScenarioMIP.
| Crossref |

Dix M, Mackallah C, Bi D, Bodman R, Marsland S, Rashid H, Woodhouse M, Druken K (2020a) CSIRO-ARCCSS ACCESS-CM2 model output prepared for CMIP6 DAMIP. Available at http://cera-www.dkrz.de/WDCC/meta/CMIP6/CMIP6.DAMIP.CSIRO-ARCCSS.ACCESS-CM2

Dix M, Mackallah C, Bi D, Bodman R, Marsland S, Rashid H, Woodhouse M, Druken K (2020b) CSIRO-ARCCSS ACCESS-CM2 model output prepared for CMIP6 RFMIP.
| Crossref |

Durack PJ, Taylor KE (2017) PCMDI AMIP SST and sea-ice boundary conditions version 1.1.3.
| Crossref |

Durack P, Taylor K, Eyring V, Ames S, Hoang T, Nadeau D, Doutriaux C, Stockhause M, Gleckler P (2018) Toward standardized data sets for climate model experimentation. Eos 99
Toward standardized data sets for climate model experimentation.Crossref | GoogleScholarGoogle Scholar |

Eyring V, Bony S, Meehl GA, Senior CA, Stevens B, Stouffer RJ, Taylor KE (2016) Overview of the Coupled Model Intercomparison Project Phase 6 (CMIP6) experimental design and organization. Geoscientific Model Development 9, 1937–1958.
Overview of the Coupled Model Intercomparison Project Phase 6 (CMIP6) experimental design and organization.Crossref | GoogleScholarGoogle Scholar |

Gillett NP, Shiogama H, Funke B, Hegerl G, Knutti R, Matthes K, Santer BD, Stone D, Tebaldi C (2016) The Detection and Attribution Model Intercomparison Project (DAMIP v1.0) contribution to CMIP6. Geoscientific Model Development 9, 3685–3697.
The Detection and Attribution Model Intercomparison Project (DAMIP v1.0) contribution to CMIP6.Crossref | GoogleScholarGoogle Scholar |

Gregory JM, Ingram WJ, Palmer MA, Jones GS, Stott PA, Thorpe RB, Lowe JA, Johns TC, Williams KD (2004) A new method for diagnosing radiative forcing and climate sensitivity. Geophysical Research Letters 31, L03205
A new method for diagnosing radiative forcing and climate sensitivity.Crossref | GoogleScholarGoogle Scholar |

Gregory JM, Bouttes N, Griffies SM, Haak H, Hurlin WJ, Jungclaus J, Kelley M, Lee WG, Marshall J, Romanou A, Saenko OA, Stammer D, Winton M (2016) The Flux-Anomaly-Forced Model Intercomparison Project (FAFMIP) contribution to CMIP6: investigation of sea-level and ocean climate change in response to CO2 forcing. Geoscientific Model Development 9, 3993–4017.
The Flux-Anomaly-Forced Model Intercomparison Project (FAFMIP) contribution to CMIP6: investigation of sea-level and ocean climate change in response to CO2 forcing.Crossref | GoogleScholarGoogle Scholar |

Griffies SM (2012) Elements of the Modular Ocean Model (MOM), GFDL Ocean Group Technical Report Number 7. (NOAA/Geophysical Fluid Dynamics Laboratory)

Griffies SM, Danabasoglu G, Durack PJ, Adcroft AJ, Balaji V, Böning CW, Chassignet EP, Curchitser E, Deshayes J, Drange H, Fox-Kemper B, Gleckler PJ, Gregory JM, Haak H, Hallberg RW, Heimbach P, Hewitt HT, Holland DM, Ilyina T, Jungclaus JH, Komuro Y, Krasting JP, Large WG, Marsland SJ, Masina S, McDougall TJ, Nurser AJG, Orr JC, Pirani A, Qiao F, Stouffer RJ, Taylor KE, Treguier AM, Tsujino H, Uotila P, Valdivieso M, Wang Q, Winton M, Yeager SG (2016) OMIP contribution to CMIP6: experimental and diagnostic protocol for the physical component of the Ocean Model Intercomparison Project. Geoscientific Model Development 9, 3231–3296.
OMIP contribution to CMIP6: experimental and diagnostic protocol for the physical component of the Ocean Model Intercomparison Project.Crossref | GoogleScholarGoogle Scholar |

Grose MR, Narsey S, Delage FP, Dowdy AJ, Bador M, Boschat G, Chung C, Kajtar JB, Rauniyar S, Freund MB, Lyu K, Rashid H, Zhang X, Wales S, Trenham C, Holbrook NJ, Cowan T, Alexander L, Arblaster JM, Power S (2020) Insights from CMIP6 for Australia’s future climate. Earth’s Future 8, e2019EF001469
Insights from CMIP6 for Australia’s future climate.Crossref | GoogleScholarGoogle Scholar |

Gutowski WJ, Giorgi F, Timbal B, Frigon A, Jacob D, Kang H-S, Raghavan K, Lee B, Lennard C, Nikulin G, O’Rourke E, Rixen M, Solman S, Stephenson T, Tangang F (2016) WCRP Coordinated Regional Downscaling Experiment (CORDEX): a diagnostic MIP for CMIP6. Geoscientific Model Development 9, 4087–4095.
WCRP Coordinated Regional Downscaling Experiment (CORDEX): a diagnostic MIP for CMIP6.Crossref | GoogleScholarGoogle Scholar |

Hardiman SC, Andrews MB, Andrews T, Bushell AC, Dunstone NJ, Dyson H, Jones GS, Knight JR, Neininger E, O’Connor FM, Ridley JK, Ringer MA, Scaife AA, Senior CA, Wood RA (2019) The impact of prescribed ozone in climate projections run with HadGEM3-GC3.1. Journal of Advances in Modeling Earth Systems 11, 3443–3453.
The impact of prescribed ozone in climate projections run with HadGEM3-GC3.1.Crossref | GoogleScholarGoogle Scholar |

Hayashida H, Kiss A, Hogg A, Hannah N, Dias FB, Brassington G, Chamberlain M, Chapman C, Dobrohotoff P, Domingues C, Duran E, England M, Fiedler R, Griffies SM, Heerdegen A, Heil P, Holmes R, Klocker A, Marsland S, Morrison A, Munroe J, Nikurashin M, Oke PR, Pilo GS, Richet O, Savita A, Spence P, Stewart KD, Ward M, Wu F, Zhang X, Mackallah C, Druken K (2021) CSIRO-COSIMA ACCESS-OM2 model output prepared for CMIP6 OMIP.
| Crossref |

Hoesly RM, Smith SJ, Feng L, Klimont Z, Janssens-Maenhout G, Pitkanen T, Seibert JJ, Vu L, Andres RJ, Bolt RM, Bond TC, Dawidowski L, Kholod N, Kurokawa J-i, Li M, Liu L, Lu Z, Moura MCP, O’Rourke PR, Zhang Q (2018) Historical (1750–2014) anthropogenic emissions of reactive gases and aerosols from the Community Emissions Data System (CEDS). Geoscientific Model Development 11, 369–408.
Historical (1750–2014) anthropogenic emissions of reactive gases and aerosols from the Community Emissions Data System (CEDS).Crossref | GoogleScholarGoogle Scholar |

Holmes R, Kiss A, Hogg A, Hannah N, Dias FB, Brassington G, Chamberlain M, Chapman C, Dobrohotoff P, Domingues C, Duran E, England M, Fiedler R, Griffies SM, Heerdegen A, Heil P, Klocker A, Marsland S, Morrison A, Munroe J, Nikurashin M, Oke PR, Pilo GS, Richet O, Savita A, Spence P, Stewart KD, Ward M, Wu F, Zhang X, Mackallah C, Druken K (2021) CSIRO-COSIMA ACCESS-OM2-025 model output prepared for CMIP6 OMIP.
| Crossref |

Holmes RM, Sohail T, Zika JD (2022) Adiabatic and diabatic signatures of ocean temperature variability. Journal of Climate 35, 1459–1477.
Adiabatic and diabatic signatures of ocean temperature variability.Crossref | GoogleScholarGoogle Scholar |

Huang Y, Wang Y-P, Ziehn T (2022) Nonlinear interactions of land carbon cycle feedbacks in Earth System Models. Global Change Biology 28, 296–306.
Nonlinear interactions of land carbon cycle feedbacks in Earth System Models.Crossref | GoogleScholarGoogle Scholar | 34687116PubMed |

Hunke EC, Lipscomb WH (2010) ‘CICE: The Los Alamos sea ice model documentation and software user’s manual, Version 4.1, LA-CC-06-012.’ (Los Alamos National Laboratory: Los Alamos, NM, USA)

Hurtt G, Chini L, Sahajpal R, Frolking S, Bodirsky BL, Calvin K, Doelman J, Fisk J, Fujimori S, Goldewijk KK, Hasegawa T, Havlik P, Heinimann A, Humpenöder F, Jungclaus J, Kaplan J, Krisztin T, Lawrence D, Lawrence P, Mertz O, Pongratz J, Popp A, Riahi K, Shevliakova E, Stehfest E, Thornton, P, van Vuuren D, Zhang X (2017) Harmonization of global land use scenarios (LUH2): Historical v2.1h 850 - 2015.
| Crossref |

Irving D, Hobbs W, Church J, Zika J (2021) A mass and energy conservation analysis of drift in the CMIP6 ensemble. Journal of Climate 34, 3157–3170.
A mass and energy conservation analysis of drift in the CMIP6 ensemble.Crossref | GoogleScholarGoogle Scholar |

Jiang L-Q, Carter BR, Feely RA, Lauvset SK, Olsen A (2019) Surface ocean pH and buffer capacity: past, present and future. Scientific Reports 9, 18624
Surface ocean pH and buffer capacity: past, present and future.Crossref | GoogleScholarGoogle Scholar | 31819102PubMed |

Jones CD, Arora V, Friedlingstein P, Bopp L, Brovkin V, Dunne J, Graven H, Hoffman F, Ilyina T, John JG, Jung M, Kawamiya M, Koven C, Pongratz J, Raddatz T, Randerson JT, Zaehle S (2016) C4MIP – the Coupled Climate–Carbon Cycle Model Intercomparison Project: experimental protocol for CMIP6. Geoscientific Model Development 9, 2853–2880.
C4MIP – the Coupled Climate–Carbon Cycle Model Intercomparison Project: experimental protocol for CMIP6.Crossref | GoogleScholarGoogle Scholar |

Jones CD, Frölicher TL, Koven C, MacDougall AH, Matthews HD, Zickfeld K, Rogelj J, Tokarska KB, Gillett NP, Ilyina T, Meinshausen M, Mengis N, Séférian R, Eby M, Burger FA (2019) The Zero Emissions Commitment Model Intercomparison Project (ZECMIP) contribution to C4MIP: quantifying committed climate changes following zero carbon emissions. Geoscientific Model Development 12, 4375–4385.
The Zero Emissions Commitment Model Intercomparison Project (ZECMIP) contribution to C4MIP: quantifying committed climate changes following zero carbon emissions.Crossref | GoogleScholarGoogle Scholar |

Jones CD, Hickman JE, Rumbold ST, Walton J, Lamboll RD, Skeie RB, Fiedler S, Forster PM, Rogelj J, Abe M, Botzet M, Calvin K, Cassou C, Cole JNS, Davini P, Deushi M, Dix M, Fyfe JC, Gillett NP, Ilyina T, Kawamiya M, Kelley M, Kharin S, Koshiro T, Li H, Mackallah C, Müller WA, Nabat P, van Noije T, Nolan P, Ohgaito R, Olivié D, Oshima N, Parodi J, Reerink TJ, Ren L, Romanou A, Séférian R, Tang Y, Timmreck C, Tjiputra J, Tourigny E, Tsigaridis K, Wang H, Wu M, Wyser K, Yang S, Yang Y, Ziehn T (2021) The climate response to emissions reductions due to COVID-19: initial results from CovidMIP. Geophysical Research Letters 48, e2020GL091883
The climate response to emissions reductions due to COVID-19: initial results from CovidMIP.Crossref | GoogleScholarGoogle Scholar | 34149115PubMed |

Juckes M, Taylor KE, Durack PJ, Lawrence B, Mizielinski MS, Pamment A, Peterschmitt J-Y, Rixen M, Sénési S (2020) The CMIP6 Data Request (DREQ, version 01.00.31). Geoscientific Model Development 13, 201–224.
The CMIP6 Data Request (DREQ, version 01.00.31).Crossref | GoogleScholarGoogle Scholar |

Kageyama M, Braconnot P, Harrison SP, Haywood AM, Jungclaus JH, Otto-Bliesner BL, Peterschmitt J-Y, Abe-Ouchi A, Albani S, Bartlein PJ, Brierley C, Crucifix M, Dolan A, Fernandez-Donado L, Fischer H, Hopcroft PO, Ivanovic RF, Lambert F, Lunt DJ, Mahowald NM, Peltier WR, Phipps SJ, Roche DM, Schmidt GA, Tarasov L, Valdes PJ, Zhang Q, Zhou T (2018) The PMIP4 contribution to CMIP6 – part 1: overview and over-arching analysis plan. Geoscientific Model Development 11, 1033–1057.
The PMIP4 contribution to CMIP6 – part 1: overview and over-arching analysis plan.Crossref | GoogleScholarGoogle Scholar |

Keller DP, Lenton A, Scott V, Vaughan NE, Bauer N, Ji D, Jones CD, Kravitz B, Muri H, Zickfeld K (2018) The Carbon Dioxide Removal Model Intercomparison Project (CDRMIP): rationale and experimental protocol for CMIP6. Geoscientific Model Development 11, 1133–1160.
The Carbon Dioxide Removal Model Intercomparison Project (CDRMIP): rationale and experimental protocol for CMIP6.Crossref | GoogleScholarGoogle Scholar |

Kiss AE, Hogg AM, Hannah N, Boeira Dias F, Brassington GB, Chamberlain MA, Chapman C, Dobrohotoff P, Domingues CM, Duran ER, England MH, Fiedler R, Griffies SM, Heerdegen A, Heil P, Holmes RM, Klocker A, Marsland SJ, Morrison AK, Munroe J, Nikurashin M, Oke PR, Pilo GS, Richet O, Savita A, Spence P, Stewart KD, Ward ML, Wu F, Zhang X (2020) ACCESS-OM2 v1.0: a global ocean–sea ice model at three resolutions. Geoscientific Model Development 13, 401–442.
ACCESS-OM2 v1.0: a global ocean–sea ice model at three resolutions.Crossref | GoogleScholarGoogle Scholar |

Kowalczyk EA, Wang YP, Law RM, Davies HL, McGregor JL, Abramowitz G (2006) The CSIRO Atmosphere Biosphere Land Exchange (CABLE) model for use in climate models and as an offline model. Marine and Atmospheric Research technical paper 13, CSIRO.

Kowalczyk EA, Stevens L, Law RM, Dix M, Wang YP, Harman IN, Haynes K, Srbinovsky J, Pak B, Ziehn T (2013) The land surface model component of ACCESS: description and impact on the simulated surface climatology. Australian Meteorological and Oceanographic Journal 63, 65–82.
The land surface model component of ACCESS: description and impact on the simulated surface climatology.Crossref | GoogleScholarGoogle Scholar |

Kwiatkowski L, Torres O, Bopp L, Aumont O, Chamberlain M, Christian JR, Dunne JP, Gehlen M, Ilyina T, John JG, Lenton A, Li H, Lovenduski NS, Orr JC, Palmieri J, Santana-Falcón Y, Schwinger J, Séférian R, Stock CA, Tagliabue A, Takano Y, Tjiputra J, Toyama K, Tsujino H, Watanabe M, Yamamoto A, Yool A, Ziehn T (2020) Twenty-first century ocean warming, acidification, deoxygenation, and upper-ocean nutrient and primary production decline from CMIP6 model projections. Biogeosciences 17, 3439–3470.
Twenty-first century ocean warming, acidification, deoxygenation, and upper-ocean nutrient and primary production decline from CMIP6 model projections.Crossref | GoogleScholarGoogle Scholar |

Law RM, Ziehn T, Matear RJ, Lenton A, Chamberlain MA, Stevens LE, Wang Y-P, Srbinovsky J, Bi D, Yan H, Vohralik PF (2017) The carbon cycle in the Australian Community Climate and Earth System Simulator (ACCESS-ESM1) – part 1: model description and pre-industrial simulation. Geoscientific Model Development 10, 2567–2590.
The carbon cycle in the Australian Community Climate and Earth System Simulator (ACCESS-ESM1) – part 1: model description and pre-industrial simulation.Crossref | GoogleScholarGoogle Scholar |

Lawrence DM, Hurtt GC, Arneth A, Brovkin V, Calvin KV, Jones AD, Jones CD, Lawrence PJ, de Noblet-Ducoudré N, Pongratz J, Seneviratne SI, Shevliakova E (2016) The Land Use Model Intercomparison Project (LUMIP) contribution to CMIP6: rationale and experimental design. Geoscientific Model Development 9, 2973–2998.
The Land Use Model Intercomparison Project (LUMIP) contribution to CMIP6: rationale and experimental design.Crossref | GoogleScholarGoogle Scholar |

Lee J, Kim J, Sun M-A, Kim B-H, Moon H, Sung HM, Kim J, Byun Y-H (2020) Evaluation of the Korea Meteorological Administration Advanced Community Earth-System model (K-ACE). Asia-Pacific Journal of Atmospheric Sciences 56, 381–395.
Evaluation of the Korea Meteorological Administration Advanced Community Earth-System model (K-ACE).Crossref | GoogleScholarGoogle Scholar |

MacDougall AH, Frölicher TL, Jones CD, Rogelj J, Matthews HD, Zickfeld K, Arora VK, Barrett NJ, Brovkin V, Burger FA, Eby M, Eliseev AV, Hajima T, Holden PB, Jeltsch-Thömmes A, Koven C, Mengis N, Menviel L, Michou M, Mokhov II, Oka A, Schwinger J, Séférian R, Shaffer G, Sokolov A, Tachiiri K, Tjiputra J, Wiltshire A, Ziehn T (2020) Is there warming in the pipeline? A multi-model analysis of the Zero Emissions Commitment from CO2. Biogeosciences 17, 2987–3016.
Is there warming in the pipeline? A multi-model analysis of the Zero Emissions Commitment from CO2.Crossref | GoogleScholarGoogle Scholar |

Mackallah C, Uhe P, Collier M (2022) ACCESS Post-Processor v4. Available at http://hdl.handle.net/102.100.100/437645?index=1

Mann GW, Carslaw KS, Spracklen DV, Ridley DA, Manktelow PT, Chipperfield MP, Pickering SJ, Johnson CE (2010) Description and evaluation of GLOMAP-mode: a modal global aerosol microphysics model for the UKCA composition-climate model. Geoscientific Model Development 3, 519–551.
Description and evaluation of GLOMAP-mode: a modal global aerosol microphysics model for the UKCA composition-climate model.Crossref | GoogleScholarGoogle Scholar |

Matthes K, Funke B, Andersson ME, Barnard L, Beer J, Charbonneau P, Clilverd MA, Dudok de Wit T, Haberreiter M, Hendry A, Jackman CH, Kretzschmar M, Kruschke T, Kunze M, Langematz U, Marsh DR, Maycock AC, Misios S, Rodger CJ, Scaife AA, Seppälä A, Shangguan M, Sinnhuber M, Tourpali K, Usoskin I, van de Kamp M, Verronen PT, Versick S (2017) Solar forcing for CMIP6 (v3.2). Geoscientific Model Development 10, 2247–2302.
Solar forcing for CMIP6 (v3.2).Crossref | GoogleScholarGoogle Scholar |

Meehl GA, Senior CA, Eyring V, Flato G, Lamarque J-F, Stouffer RJ, Taylor KE, Schlund M (2020) Context for interpreting equilibrium climate sensitivity and transient climate response from the CMIP6 Earth system models. Science Advances 6, eaba1981
Context for interpreting equilibrium climate sensitivity and transient climate response from the CMIP6 Earth system models.Crossref | GoogleScholarGoogle Scholar | 32637602PubMed |

Meinshausen M, Vogel E (2016) input4MIPs.UoM.GHGConcentrations.CMIP.UoM-CMIP-1-2-0.
| Crossref |

Meinshausen M, Vogel E, Nauels A, Lorbacher K, Meinshausen N, Etheridge DM, Fraser PJ, Montzka SA, Rayner PJ, Trudinger CM, Krummel PB, Beyerle U, Canadell JG, Daniel JS, Enting IG, Law RM, Lunder CR, O’Doherty S, Prinn RG, Reimann S, Rubino M, Velders GJM, Vollmer MK, Wang RHJ, Weiss R (2017) Historical greenhouse gas concentrations for climate modelling (CMIP6). Geoscientific Model Development 10, 2057–2116.
Historical greenhouse gas concentrations for climate modelling (CMIP6).Crossref | GoogleScholarGoogle Scholar |

Meinshausen M, Nicholls ZRJ, Lewis J, Gidden MJ, Vogel E, Freund M, Beyerle U, Gessner C, Nauels A, Bauer N, Canadell JG, Daniel JS, John A, Krummel PB, Luderer G, Meinshausen N, Montzka SA, Rayner PJ, Reimann S, Smith SJ, van den Berg M, Velders GJM, Vollmer MK, Wang RHJ (2020) The shared socio-economic pathway (SSP) greenhouse gas concentrations and their extensions to 2500. Geoscientific Model Development 13, 3571–3605.
The shared socio-economic pathway (SSP) greenhouse gas concentrations and their extensions to 2500.Crossref | GoogleScholarGoogle Scholar |

Met Office (2020) Iris: a Python library for analysing and visualising meteorological and oceanographic data sets, Exeter, Devon, v3 edn. Available at http://scitools.org.uk/

Morgenstern O, Hegglin MI, Rozanov E, O’Connor FM, Abraham NL, Akiyoshi H, Archibald AT, Bekki S, Butchart N, Chipperfield MP, Deushi M, Dhomse SS, Garcia RR, Hardiman SC, Horowitz LW, Jöckel P, Josse B, Kinnison D, Lin M, Mancini E, Manyin ME, Marchand M, Marécal V, Michou M, Oman LD, Pitari G, Plummer DA, Revell LE, Saint-Martin D, Schofield R, Stenke A, Stone K, Sudo K, Tanaka TY, Tilmes S, Yamashita Y, Yoshida K, Zeng G (2017) Review of the global models used within phase 1 of the Chemistry–Climate Model Initiative (CCMI). Geoscientific Model Development 10, 639–671.
Review of the global models used within phase 1 of the Chemistry–Climate Model Initiative (CCMI).Crossref | GoogleScholarGoogle Scholar |

Nadeau D, Doutriaux C, Bradshaw T, Kettleborough J, Weigel T, Hogan E, Durack PJ (2016) PCMDI/cmor: CMOR version 3.2.0.
| Crossref |

Oke PR, Griffin DA, Schiller A, Matear RJ, Fiedler R, Mansbridge J, Lenton A, Cahill M, Chamberlain MA, Ridgway K (2013) Evaluation of a near-global eddy-resolving ocean model. Geoscientific Model Development 6, 591–615.
Evaluation of a near-global eddy-resolving ocean model.Crossref | GoogleScholarGoogle Scholar |

O’Neill BC, Tebaldi C, van Vuuren DP, Eyring V, Friedlingstein P, Hurtt G, Knutti R, Kriegler E, Lamarque J-F, Lowe J, Meehl GA, Moss R, Riahi K, Sanderson BM (2016) The Scenario Model Intercomparison Project (ScenarioMIP) for CMIP6. Geoscientific Model Development 9, 3461–3482.
The Scenario Model Intercomparison Project (ScenarioMIP) for CMIP6.Crossref | GoogleScholarGoogle Scholar |

Orr JC, Najjar RG, Aumont O, Bopp L, Bullister JL, Danabasoglu G, Doney SC, Dunne JP, Dutay J-C, Graven H, Griffies SM, John JG, Joos F, Levin I, Lindsay K, Matear RJ, McKinley GA, Mouchet A, Oschlies A, Romanou A, Schlitzer R, Tagliabue A, Tanhua T, Yool A (2017) Biogeochemical protocols and diagnostics for the CMIP6 Ocean Model Intercomparison Project (OMIP). Geoscientific Model Development 10, 2169–2199.
Biogeochemical protocols and diagnostics for the CMIP6 Ocean Model Intercomparison Project (OMIP).Crossref | GoogleScholarGoogle Scholar |

Otto-Bliesner BL, Braconnot P, Harrison SP, Lunt DJ, Abe-Ouchi A, Albani S, Bartlein PJ, Capron E, Carlson AE, Dutton A, Fischer H, Goelzer H, Govin A, Haywood A, Joos F, LeGrande AN, Lipscomb WH, Lohmann G, Mahowald N, Nehrbass-Ahles C, Pausata FSR, Peterschmitt J-Y, Phipps SJ, Renssen H, Zhang Q (2017) The PMIP4 contribution to CMIP6 – part 2: two interglacials, scientific objective and experimental design for Holocene and Last Interglacial simulations. Geoscientific Model Development 10, 3979–4003.
The PMIP4 contribution to CMIP6 – part 2: two interglacials, scientific objective and experimental design for Holocene and Last Interglacial simulations.Crossref | GoogleScholarGoogle Scholar |

Pincus R, Forster PM, Stevens B (2016) The Radiative Forcing Model Intercomparison Project (RFMIP): experimental protocol for CMIP6. Geoscientific Model Development 9, 3447–3460.
The Radiative Forcing Model Intercomparison Project (RFMIP): experimental protocol for CMIP6.Crossref | GoogleScholarGoogle Scholar |

Rashid HA, Sullivan A, Dix M, Bi D, Mackallah C, Ziehn T, Dobrohotoff P, O’Farrell S, Harman IN, Bodman R, Marsland S (2022) Evaluation of climate variability and change in ACCESS historical simulations for CMIP6. Journal of Southern Hemisphere Earth Systems Science
Evaluation of climate variability and change in ACCESS historical simulations for CMIP6.Crossref | GoogleScholarGoogle Scholar | [Published online xx July 2022].

Rugenstein M, Bloch-Johnson J, Abe-Ouchi A, Andrews T, Beyerle U, Cao L, Chadha T, Danabasoglu G, Dufresne J-L, Duan L, Foujols M-A, Frölicher T, Geoffroy O, Gregory J, Knutti R, Li C, Marzocchi A, Mauritsen T, Menary M, Moyer E, Nazarenko L, Paynter D, Saint-Martin D, Schmidt GA, Yamamoto A, Yang S (2019) LongRunMIP: motivation and design for a large collection of millennial-length AOGCM simulations. Bulletin of the American Meteorological Society 100, 2551–2570.
LongRunMIP: motivation and design for a large collection of millennial-length AOGCM simulations.Crossref | GoogleScholarGoogle Scholar |

Savita A, Marsland S, Dix M, Bi D, Dobrohotoff P, Fiedler R, Mackallah C, Sullivan A, Dias FB, Domingues C, Hannah N, Heerdegen A, Hogg A, Druken K (2019) CSIRO-ARCCSS ACCESS-CM2 model output prepared for CMIP6 FAFMIP.
| Crossref |

Sellar AA, Walton J, Jones CG, Wood R, Abraham NL, Andrejczuk M, Andrews MB, Andrews T, Archibald AT, de Mora L, Dyson H, Elkington M, Ellis R, Florek P, Good P, Gohar L, Haddad S, Hardiman SC, Hogan E, Iwi A, Jones CD, Johnson B, Kelley DI, Kettleborough J, Knight JR, Köhler MO, Kuhlbrodt T, Liddicoat S, Linova-Pavlova I, Mizielinski MS, Morgenstern O, Mulcahy J, Neininger E, O’Connor FM, Petrie R, Ridley J, Rioual J-C, Roberts M, Robertson E, Rumbold S, Seddon J, Shepherd H, Shim S, Stephens A, Teixiera JC, Tang Y, Williams J, Wiltshire A, Griffiths PT (2020) Implementation of UK earth system models for CMIP6. Journal of Advances in Modeling Earth Systems 12, e2019MS001946
Implementation of UK earth system models for CMIP6.Crossref | GoogleScholarGoogle Scholar |

Senior CA, Jones CG, Wood RA, Sellar A, Belcher S, Klein-Tank A, Sutton R, Walton J, Lawrence B, Andrews T, Mulcahy JP (2020) UK community earth system modeling for CMIP6. Journal of Advances in Modeling Earth Systems 12, e2019MS002004
UK community earth system modeling for CMIP6.Crossref | GoogleScholarGoogle Scholar | 33042388PubMed |

Sherwood SC, Webb MJ, Annan JD, Armour KC, Forster PM, Hargreaves JC, Hegerl G, Klein SA, Marvel KD, Rohling EJ, Watanabe M, Andrews T, Braconnot P, Bretherton CS, Foster GL, Hausfather Z, von der Heydt AS, Knutti R, Mauritsen T, Norris JR, Proistosescu C, Rugenstein M, Schmidt GA, Tokarska KB, Zelinka MD (2020) An assessment of earth’s climate sensitivity using multiple lines of evidence. Reviews of Geophysics 58, e2019RG000678
An assessment of earth’s climate sensitivity using multiple lines of evidence.Crossref | GoogleScholarGoogle Scholar | 33015673PubMed |

Smith CJ, Kramer RJ, Myhre G, Alterskjær K, Collins W, Sima A, Boucher O, Dufresne J-L, Nabat P, Michou M, Yukimoto S, Cole J, Paynter D, Shiogama H, O’Connor FM, Robertson E, Wiltshire A, Andrews T, Hannay C, Miller R, Nazarenko L, Kirkevåg A, Olivié D, Fiedler S, Lewinschal A, Mackallah C, Dix M, Pincus R, Forster PM (2020) Effective radiative forcing and adjustments in CMIP6 models. Atmospheric Chemistry and Physics 20, 9591–9618.
Effective radiative forcing and adjustments in CMIP6 models.Crossref | GoogleScholarGoogle Scholar |

Storkey D, Blaker AT, Mathiot P, Megann A, Aksenov Y, Blockley EW, Calvert D, Graham T, Hewitt HT, Hyder P, Kuhlbrodt T, Rae JGL, Sinha B (2018) UK Global Ocean GO6 and GO7: a traceable hierarchy of model resolutions. Geoscientific Model Development 11, 3187–3213.
UK Global Ocean GO6 and GO7: a traceable hierarchy of model resolutions.Crossref | GoogleScholarGoogle Scholar |

Taylor K, Williamson D, Zwiers F (2000) AMIP II sea surface temperature and sea ice concentration boundary conditions. PCMDI Report, Vol. 60.

Thomason LW, Ernest N, Millán L, Rieger L, Bourassa A, Vernier J-P, Manney G, Luo B, Arfeuille F, Peter T (2018) A global space-based stratospheric aerosol climatology: 1979-2016. Earth System Science Data 10, 469–492.
A global space-based stratospheric aerosol climatology: 1979-2016.Crossref | GoogleScholarGoogle Scholar |

Toohey M, Stevens B, Schmidt H, Timmreck C (2016) Easy Volcanic Aerosol (EVA v1.0): an idealized forcing generator for climate simulations. Geoscientific Model Development 9, 4049–4070.
Easy Volcanic Aerosol (EVA v1.0): an idealized forcing generator for climate simulations.Crossref | GoogleScholarGoogle Scholar |

Tsujino H, Urakawa S, Nakano H, Small RJ, Kim WM, Yeager SG, Danabasoglu G, Suzuki T, Bamber JL, Bentsen M, Böning CW, Bozec A, Chassignet EP, Curchitser E, Boeira Dias F, Durack PJ, Griffies SM, Harada Y, Ilicak M, Josey SA, Kobayashi C, Kobayashi S, Komuro Y, Large WG, Le Sommer J, Marsland SJ, Masina S, Scheinert M, Tomita H, Valdivieso M, Yamazaki D (2018) JRA-55 based surface dataset for driving ocean–sea-ice models (JRA55-do). Ocean Modelling 130, 79–139.
JRA-55 based surface dataset for driving ocean–sea-ice models (JRA55-do).Crossref | GoogleScholarGoogle Scholar |

Tsujino H, Urakawa LS, Griffies SM, Danabasoglu G, Adcroft AJ, Amaral AE, Arsouze T, Bentsen M, Bernardello R, Böning CW, Bozec A, Chassignet EP, Danilov S, Dussin R, Exarchou E, Fogli PG, Fox-Kemper B, Guo C, Ilicak M, Iovino D, Kim WM, Koldunov N, Lapin V, Li Y, Lin P, Lindsay K, Liu H, Long MC, Komuro Y, Marsland SJ, Masina S, Nummelin A, Rieck JK, Ruprich-Robert Y, Scheinert M, Sicardi V, Sidorenko D, Suzuki T, Tatebe H, Wang Q, Yeager SG, Yu Z (2020) Evaluation of global ocean–sea-ice model simulations based on the experimental protocols of the Ocean Model Intercomparison Project phase 2 (OMIP-2). Geoscientific Model Development 13, 3643–3708.
Evaluation of global ocean–sea-ice model simulations based on the experimental protocols of the Ocean Model Intercomparison Project phase 2 (OMIP-2).Crossref | GoogleScholarGoogle Scholar |

Uhe P, Hume T, Collier M (2012) ACCESS Post-Processor Version 1.0, CAWCR Technical Report Number 058. Available at https://www.cawcr.gov.au/technical-reports/CTR_058.pdf

van Marle MJE, Kloster S, Magi BI, Marlon JR, Daniau A-L, Field RD, Arneth A, Forrest M, Hantson S, Kehrwald NM, Knorr W, Lasslop G, Li F, Mangeon S, Yue C, Kaiser JW, van der Werf GR (2017) Historic global biomass burning emissions for CMIP6 (BB4CMIP) based on merging satellite observations with proxies and fire models (1750-2015). Geoscientific Model Development 10, 3329–3357.
Historic global biomass burning emissions for CMIP6 (BB4CMIP) based on merging satellite observations with proxies and fire models (1750-2015).Crossref | GoogleScholarGoogle Scholar |

Walters D, Baran AJ, Boutle I, Brooks M, Earnshaw P, Edwards J, Furtado K, Hill P, Lock A, Manners J, Morcrette C, Mulcahy J, Sanchez C, Smith C, Stratton R, Tennant W, Tomassini L, Van Weverberg K, Vosper S, Willett M, Browse J, Bushell A, Carslaw K, Dalvi M, Essery R, Gedney N, Hardiman S, Johnson B, Johnson C, Jones A, Jones C, Mann G, Milton S, Rumbold H, Sellar A, Ujiie M, Whitall M, Williams K, Zerroukat M (2019) The Met Office Unified Model Global Atmosphere 7.0/7.1 and JULES Global Land 7.0 configurations. Geoscientific Model Development 12, 1909–1963.
The Met Office Unified Model Global Atmosphere 7.0/7.1 and JULES Global Land 7.0 configurations.Crossref | GoogleScholarGoogle Scholar |

Wang YP, Law RM, Pak B (2010) A global model of carbon, nitrogen and phosphorus cycles for the terrestrial biosphere. Biogeosciences 7, 2261–2282.
A global model of carbon, nitrogen and phosphorus cycles for the terrestrial biosphere.Crossref | GoogleScholarGoogle Scholar |

Watterson IG, Keane RJ, Dix M, Ziehn T, Andrews T, Tang Y (2021) Analysis of CMIP6 atmospheric moisture fluxes and the implications for projections of future change in mean and heavy rainfall. International Journal of Climatology 41, E1417–E1434.
Analysis of CMIP6 atmospheric moisture fluxes and the implications for projections of future change in mean and heavy rainfall.Crossref | GoogleScholarGoogle Scholar |

Woodward S (2011) Mineral dust in HadGEM2. Technical note 87, Hadley Centre.

Yeung N, Menviel L, Meissner K, Ziehn T, Chamberlain M, Mackallah C, Druken K, Ridzwan SM (2019) CSIRO ACCESS-ESM1.5 model output prepared for CMIP6 PMIP.
| Crossref |

Yeung NK-H, Menviel L, Meissner KJ, Taschetto AS, Ziehn T, Chamberlain M (2021) Land–sea temperature contrasts at the Last Interglacial and their impact on the hydrological cycle. Climate of the Past 17, 869–885.
Land–sea temperature contrasts at the Last Interglacial and their impact on the hydrological cycle.Crossref | GoogleScholarGoogle Scholar |

Zelinka MD, Myers TA, McCoy DT, Po-Chedley S, Caldwell PM, Ceppi P, Klein SA, Taylor KE (2020) Causes of higher climate sensitivity in CMIP6 Models. Geophysical Research Letters 47, e2019GL085782
Causes of higher climate sensitivity in CMIP6 Models.Crossref | GoogleScholarGoogle Scholar |

Zhang Q, Pitman AJ, Wang YP, Dai YJ, Lawrence PJ (2013) The impact of nitrogen and phosphorous limitation on the estimated terrestrial carbon balance and warming of land use change over the last 156 yr. Earth System Dynamics 4, 333–345.
The impact of nitrogen and phosphorous limitation on the estimated terrestrial carbon balance and warming of land use change over the last 156 yr.Crossref | GoogleScholarGoogle Scholar |

Ziehn T, Chamberlain M, Lenton A, Law R, Bodman R, Dix M, Mackallah C, Druken K, Ridzwan SM (2019a) CSIRO ACCESS-ESM1.5 model output prepared for CMIP6 C4MIP.
| Crossref |

Ziehn T, Chamberlain M, Lenton A, Law R, Bodman R, Dix M, Mackallah C, Druken K, Ridzwan SM (2019b) CSIRO ACCESS-ESM1.5 model output prepared for CMIP6 CDRMIP.
| Crossref |

Ziehn T, Chamberlain M, Lenton A, Law R, Bodman R, Dix M, Wang Y, Dobrohotoff P, Srbinovsky J, Stevens L, Vohralik P, Mackallah C, Sullivan A, O’Farrell S, Druken K (2019c) CSIRO ACCESS-ESM1.5 model output prepared for CMIP6 CMIP.
| Crossref |

Ziehn T, Chamberlain M, Lenton A, Law R, Bodman R, Dix M, Wang Y, Dobrohotoff P, Srbinovsky J, Stevens L, Vohralik P, Mackallah C, Sullivan A, O’Farrell S, Druken K (2019d) CSIRO ACCESS-ESM1.5 model output prepared for CMIP6 ScenarioMIP.
| Crossref |

Ziehn T, Chamberlain MA, Law RM, Lenton A, Bodman RW, Dix M, Stevens L, Wang YP, Srbinovsky J (2020a) The Australian Earth System Model: ACCESS-ESM1.5. Journal of Southern Hemisphere Earth Systems Science 70, 193–214.
The Australian Earth System Model: ACCESS-ESM1.5.Crossref | GoogleScholarGoogle Scholar |

Ziehn T, Dix M, Mackallah C, Chamberlain M, Lenton A, Law R, Druken K, Ridzwan SM (2020b) CSIRO ACCESS-ESM1.5 model output prepared for CMIP6 DAMIP.
| Crossref |

Ziehn T, Dix M, Mackallah C, Chamberlain M, Lenton A, Law R, Druken K, Ridzwan SM (2020c) CSIRO ACCESS-ESM1.5 model output prepared for CMIP6 RFMIP.
| Crossref |

Ziehn T, Lenton A, Law R (2020d) An assessment of land-based climate and carbon reversibility in the Australian Community Climate and Earth System Simulator. Mitigation and Adaptation Strategies for Global Change 25, 713–731.
An assessment of land-based climate and carbon reversibility in the Australian Community Climate and Earth System Simulator.Crossref | GoogleScholarGoogle Scholar |

Ziehn T, Dix M, Mackallah C, Chamberlain M, Lenton A, Law R, Druken K, Ridzwan SM (2021a) CSIRO ACCESS-ESM1.5 model output prepared for CMIP6 LUMIP. Available at http://cera-www.dkrz.de/WDCC/meta/CMIP6/CMIP6.LUMIP.CSIRO.ACCESS-ESM1-5

Ziehn T, Wang Y-P, Huang Y (2021b) Land carbon-concentration and carbon-climate feedbacks are significantly reduced by nitrogen and phosphorus limitation. Environmental Research Letters 16, 074043
Land carbon-concentration and carbon-climate feedbacks are significantly reduced by nitrogen and phosphorus limitation.Crossref | GoogleScholarGoogle Scholar |




1 https://www.wcrp-climate.org/modelling-wgcm-mip-catalogue/modelling-wgcm-cmip6-endorsed-mips

2 That is, global climate models that simulate the carbon cycle, such as ACCESS-ESM1.5.

3 https://github.com/WCRP-CMIP/CMIP6_CVs

4 http://www.longrunmip.org/

5 https://large-ensemble.github.io/index

6 https://esgf-node.llnl.gov/projects/input4mips

7 http://www.globalchange.umd.edu/ceds/ceds-cmip6-data

8 see errata: https://errata.es-doc.org/static/view.html?uid=5fda6ba1-3b14-38a1-94b9-22a09ad7e92e

9 see https://help.ceda.ac.uk/article/4424-pp-binary-format

10 https://cdms.readthedocs.io/en/latest/cdms2.html

11 https://cmor.llnl.gov

12 see https://goo.gl/v1drZl ‘CMIP6 Global Attributes, DRS, Filenames, Directory Structure, and CV’s’.

13 https://pcmdi.llnl.gov/CMIP6/Guide/modelers.html

14 https://github.com/PCMDI/cmip6-cmor-tables

15 ACCESS-ESM1.5 is spelled as ACCESS-ESM1-5 on the ESGF in order to follow CMIP6 conventions concerningmodel naming.

16 http://esgf-ui.cmcc.it/esgf-dashboard-ui/data-archiveCMIP6.html

17 https://errata.es-doc.org/static/index.html

18 https://metomi.github.io/rose

19 https://cylc.github.io/cylc

20 http://esgf-ui.cmcc.it/esgf-dashboard-ui/cmip6.html

21 https://cmip-publications.llnl.gov

22 https://scholar.google.com/citations?hl=en user=inVqu_4AAAAJ

23 https://cordex.org/

24 https://cowclip.org/

25 https://www.access-nri.org.au/